24.08.2013 Views

Synthetic Molecular Motors and Mechanical Machines - Engineering ...

Synthetic Molecular Motors and Mechanical Machines - Engineering ...

Synthetic Molecular Motors and Mechanical Machines - Engineering ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Reviews<br />

<strong>Molecular</strong> Devices<br />

DOI: 10.1002/anie.200504313<br />

<strong>Synthetic</strong> <strong>Molecular</strong> <strong>Motors</strong> <strong>and</strong> <strong>Mechanical</strong> <strong>Machines</strong><br />

Euan R. Kay, David A. Leigh,* <strong>and</strong> Francesco Zerbetto*<br />

Keywords:<br />

molecular devices · nanotechnology ·<br />

noncovalent interactions ·<br />

supramolecular chemistry<br />

Angew<strong>and</strong>te<br />

Chemie<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

72 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

The widespread use of controlled molecular-level motion in key<br />

natural processes suggests that great rewards could come from<br />

bridging the gap between the present generation of synthetic<br />

molecular systems, which by <strong>and</strong> large rely upon electronic <strong>and</strong><br />

chemical effects to carry out their functions, <strong>and</strong> the machines of<br />

the macroscopic world, which utilize the synchronized movements<br />

of smaller parts to perform specific tasks. This is a scientific<br />

area of great contemporary interest <strong>and</strong> extraordinary recent<br />

growth, yet the notion of molecular-level machines dates back to a<br />

time when the ideas surrounding the statistical nature of matter<br />

<strong>and</strong> the laws of thermodynamics were first being formulated. Here<br />

we outline the exciting successes in taming molecular-level<br />

movement thus far, the underlying principles that all experimental<br />

designs must follow, <strong>and</strong> the early progress made towards utilizing<br />

synthetic molecular structures to perform tasks using mechanical<br />

motion. We also highlight some of the issues <strong>and</strong> challenges that<br />

still need to be overcome.<br />

1. Design Principles for <strong>Molecular</strong>-Level <strong>Motors</strong> <strong>and</strong><br />

<strong>Machines</strong><br />

In recent years chemists have demonstrated imagination<br />

<strong>and</strong> considerable skill in the design <strong>and</strong> construction of<br />

synthetic molecular systems in which positional displacements<br />

of submolecular components result from moving<br />

downhill in terms of energy. [1] But what are the structural<br />

features necessary for molecules to use chemical energy to<br />

repetitively do mechanical work? How can we make a<br />

molecular machine that pumps ions to reverse a concentration<br />

gradient,for example,or moves itself uphill in terms of<br />

energy? How can we make nanoscale structures that traverse<br />

a predefined path across a surface or down a track by<br />

responding to the nature of their environment so as to change<br />

direction? How can we make a synthetic molecular motor<br />

that rotates against an applied torque? Artificial compounds<br />

that can do such things have yet to be realized. Moreover,<strong>and</strong><br />

perhaps more surprisingly given that nature has developed<br />

such machines <strong>and</strong> refined them to a high degree of efficiency,<br />

the literature is still largely bereft of the fundamental<br />

guidelines necessary to invent them. In this Review we try<br />

to address this deficiency by using theory to outline,develop,<br />

<strong>and</strong> present the underlying mechanisms <strong>and</strong> principles<br />

required for future generation synthetic molecular-level<br />

machine systems alongside the current experimental state of<br />

the art.<br />

Perhaps inevitably in an emerging field,there is not even a<br />

clear consensus as to what constitutes a molecular machine<br />

<strong>and</strong> what differentiates them from other molecular devices.<br />

Initially,the categorization of molecules as machines by<br />

chemists was purely iconic—the structures “looked” like<br />

pieces of machinery—or they were so-called because they<br />

carried out a function that in the macroscopic world would<br />

require a machine to perform it. Many of the chemical<br />

systems first likened to pistons <strong>and</strong> other machines were<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim<br />

From the Contents<br />

1. Design Principles for <strong>Molecular</strong>-Level<br />

<strong>Motors</strong> <strong>and</strong> <strong>Machines</strong> 73<br />

2. Controlling Motion in Covalently<br />

Bonded <strong>Molecular</strong> Systems 85<br />

3. Controlling Motion in<br />

Supramolecular Systems 99<br />

4. Controlling Motion in <strong>Mechanical</strong>ly<br />

Bonded <strong>Molecular</strong> Systems 101<br />

5. <strong>Molecular</strong>-Level Motion Driven by<br />

External Fields 128<br />

6. Self-Propelled “Nanostructures” 132<br />

7. <strong>Molecular</strong>-Level Motion Driven by<br />

Atomically Sharp Tools 134<br />

8. From Laboratory to Technology:<br />

Towards Useful <strong>Molecular</strong> <strong>Machines</strong> 139<br />

9. Artificial Biomolecular <strong>Machines</strong> 157<br />

10. Conclusions <strong>and</strong> Outlook 163<br />

simply host–guest complexes in which the binding could be<br />

switched “on” or “off” by external stimuli such as light. [1]<br />

Whilst such studies were instrumental in popularizing the<br />

notion of molecular-level machines amongst chemists,it is fair<br />

to say (with considerable hindsight [2] ) that the effects of scale<br />

tell us that supramolecular decomplexation events have little<br />

in common with the motion or function of a piston (the<br />

analogy is somewhat better applied to shuttling within a<br />

rotaxane architecture because the components are always<br />

kinetically associated,but the implication of imparting<br />

momentum is still unfortunate). Similarly,a photosensitizer<br />

is not phenomenologically related to a “light-fueled motor”.<br />

Here we choose to differentiate machines from other devices<br />

on the basis that the etymology <strong>and</strong> meaning of “machine”<br />

generally implies mechanical movement—that is,a net<br />

nuclear displacement in the molecular world—which causes<br />

something useful to happen. Thus,for our purposes,“molec-<br />

[*] E. R. Kay, Prof. D. A. Leigh<br />

School of Chemistry<br />

University of Edinburgh<br />

The King’s Buildings, West Mains Road, Edinburgh EH93JJ (UK)<br />

Fax: (+ 44)131-650-6453<br />

E-mail: david.leigh@ed.ac.uk<br />

Prof. F. Zerbetto<br />

Dipartimento di Chimica “G. Ciamician”<br />

Università di Bologna<br />

v. F. Selmi 2, 40126 Bologna (Italy)<br />

Fax: (+ 39)051-209-9456<br />

E-mail: francesco.zerbetto@unibo.it<br />

Angew<strong>and</strong>te<br />

Chemie<br />

73


Reviews<br />

ular machines” are a defined subset of “molecular devices”<br />

(functional molecular systems) in which some stimulus<br />

triggers the controlled,large amplitude or directional<br />

mechanical motion of one component relative to another<br />

(or of a substrate relative to the machine) which results in a<br />

net task being performed. Accordingly,we shall not discuss<br />

here supramolecular complexes in which the components are<br />

able to exchange with others in the bulk,or systems that<br />

function solely through changes in their electronic structure.<br />

Rather we will limit our scope to mechanical molecular-level<br />

devices,systems that feature—or attempt to feature—a<br />

degree of control over the motion of the components or<br />

substrates. The examples are chosen to help demonstrate the<br />

requirements for performing mechanical tasks at the molecular<br />

level.<br />

1.1. <strong>Molecular</strong>-Level <strong>Machines</strong> <strong>and</strong> the Language Used to<br />

Describe Them<br />

Language—especially scientific language—has to be suitably<br />

defined <strong>and</strong> correctly used to accurately convey concepts<br />

Euan Kay was born in Lanarkshire (Scotl<strong>and</strong>)<br />

<strong>and</strong> educated at Hutchesons’ Grammar<br />

School. He received his MChem from<br />

the University of Edinburgh in 2002. He is<br />

currently a Carnegie Trust Scholar working<br />

towards a PhD in the group of David Leigh<br />

on developing <strong>and</strong> demonstrating mechanisms<br />

for the control of molecular-level<br />

motion in mechanically interlocked<br />

molecules.<br />

David Leigh was born in Birmingham (UK)<br />

<strong>and</strong> obtained his BSc <strong>and</strong> PhD from the<br />

University of Sheffield. After postdoctoral<br />

research at the NRC of Canada in Ottawa<br />

(1987–1989) he returned to the UK as a<br />

lecturer at UMIST. In 1998 he moved to the<br />

University of Warwick, <strong>and</strong> in 2001 he took<br />

up the Forbes Chair of Organic Chemistry at<br />

the University of Edinburgh. He holds both<br />

an EPSRC Senior Research Fellowship <strong>and</strong> a<br />

RoyalSociety-Wolfson Research Merit<br />

Award, <strong>and</strong> is a Fellow of the Royal Society<br />

of Edinburgh, Scotl<strong>and</strong>’s National Academy<br />

of Science <strong>and</strong> Letters. His research interests include molecular-level<br />

motors <strong>and</strong> machines.<br />

Francesco Zerbetto received his PhD from<br />

the University of Bologna in 1986. From<br />

1986 to 1990 he was a postdoctoralresearch<br />

associate at the NationalResearch Council<br />

of Canada in Ottawa. He is currently Professor<br />

of PhysicalChemistry at the University of<br />

Bologna. His current research interests focus<br />

on the simulation of large systems that<br />

range from mechanically interlocked molecules,<br />

to fullerenes <strong>and</strong> nanotubes, to the<br />

interface between inorganic surfaces <strong>and</strong><br />

organic materials. He is the author of more<br />

than 200 papers.<br />

in a field. Nowhere is the need for accurate scientific language<br />

more apparent than in the discussion of the ideas <strong>and</strong><br />

mechanisms by which nanoscale machines could—<strong>and</strong> do—<br />

operate. [3] Much of the terminology used to describe molecular-level<br />

machines has its origins in the observations of<br />

physicists <strong>and</strong> biologists,but their findings <strong>and</strong> descriptions<br />

have at times been misunderstood or under-appreciated by<br />

chemists. Similarly,the chemistry of molecular systems can<br />

sometimes be overlooked by other disciplines. [4]<br />

As mechanism replaces imagery as the driving force<br />

behind advances in this field,it may be helpful for the<br />

terminology used to discuss molecular-level machine systems<br />

to become more phenomenologically based. When describing<br />

molecular behavior scientifically,the st<strong>and</strong>ard dictionary<br />

definitions meant for everyday use are not always appropriate<br />

for a regime that the definitions were never intended to cover.<br />

The difference between a “motor” <strong>and</strong> a “switch” as basic<br />

molecular machine types,for example,is significant because<br />

“motor” <strong>and</strong> “switch” become descriptors of very different<br />

types of behavior at molecular length scales,not simply iconic<br />

images (Figure 1). A “switch” influences a system as a<br />

function of state whereas a “motor” influences a system as a<br />

function of the trajectory of its components or the substrate.<br />

Returning a molecular-level switch to its original position<br />

undoes any mechanical effect it has on an external system<br />

(naturally,the molecules heat up their surroundings as the<br />

energy from the switching stimulus is dissipated); when the<br />

components of a rotary motor return to their original position<br />

through a different pathway to the one they left by (namely,a<br />

3608 directional rotation),a physical task performed by the<br />

machine is not inherently undone (for example,the rotating<br />

components could be used to wind up a polymer chain).<br />

This difference is profound <strong>and</strong> the terms really should<br />

not be used interchangeably as sometimes happens in the<br />

chemistry (but not physics [5] or biology) literature. That is not<br />

to say that molecular switches cannot use chemical energy to<br />

do mechanical work. They can,but it is undone by resetting<br />

the switch to its original state. The key reason why this point is<br />

important is that switches cannot use chemical energy to<br />

repetitively <strong>and</strong> progressively drive a system away from<br />

equilibrium,whereas a motor can. [6] Other molecular machine<br />

types,also differentiated on the basis of their fundamental<br />

behavior,can also be identified (see Section 4.4). It is an<br />

accurate reflection of the current state of the art that the vast<br />

majority of the molecular-level machines discussed in this<br />

Review are switches,not motors. Similarly,thus far only<br />

“influence-as-a-function-of-state” rather then “influence-asa-function-of-trajectory”<br />

molecular-level machines have been<br />

developed to the level that they can be exploited to perform<br />

tasks in the outside world.<br />

1.2. The Effects of Scale<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

The path towards synthetic molecular machines can be<br />

traced back nearly two centuries to the observation of effects<br />

that pointed directly to the r<strong>and</strong>om motion experienced by all<br />

molecular-scale objects. In 1827,the Scottish botanist Robert<br />

Brown noted through his microscope the incessant,haphaz-<br />

74 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 1. The fundamental difference between a “switch” <strong>and</strong> a<br />

“motor” at the molecular level. Both translational <strong>and</strong> rotary switches<br />

influence a system as a function of the switch state. They switch<br />

between two or more, often equilibrium, states. <strong>Motors</strong>, however,<br />

influence a system as a function of the trajectory of their components<br />

or a substrate. <strong>Motors</strong> function repetitively <strong>and</strong> progressively on a<br />

system; the affect of a switch is undone by resetting the machine.<br />

a) Rotary switch. b) Rotary motor. c) Translational switch. d) Translational<br />

motor or pump.<br />

ard motion of tiny particles within translucent pollen grains<br />

suspended in water. [7] An explanation of the phenomenon—<br />

now known as Brownian motion or movement—was provided<br />

by Einstein in one [8] of his three celebrated papers of 1905 <strong>and</strong><br />

was proven experimentally [9] by Perrin over the next<br />

decade. [10] Scientists have been fascinated by the implications<br />

of the stochastic nature of molecular-level motion ever since.<br />

The r<strong>and</strong>om thermal fluctuations experienced by molecules<br />

dominate mechanical behavior in the molecular world. Even<br />

the most efficient nanoscale machines are swamped by its<br />

effect. A typical motor protein consumes ATP fuel at a rate of<br />

100–1000 molecules every second,which corresponds to a<br />

maximum possible power output in the region of 10 16 to<br />

10 17 W per molecule. [11] When compared with the r<strong>and</strong>om<br />

environmental buffeting of approximately 10 8 W experienced<br />

by molecules in solution at room temperature,it seems<br />

remarkable that any form of controlled motion is possible. [12]<br />

When designing molecular machines it is important to<br />

remember that the presence of Brownian motion is a<br />

consequence of scale,not of the nature of the surroundings.<br />

It cannot be avoided by putting a molecular-level structure in<br />

a near-vacuum,for example. Although there would be few<br />

r<strong>and</strong>om collisions to set such a Brownian particle in motion,<br />

there would be little viscosity to slow it down. These effects<br />

always cancel each other out,<strong>and</strong> as long as a temperature can<br />

be defined for an object it will undergo Brownian motion<br />

appropriate to that temperature (which determines the<br />

kinetic energy of the particle) <strong>and</strong> only the mean-free path<br />

between collisions is affected by the concentration of<br />

particles. In the absence of any other molecules,heat would<br />

still be transmitted from the hot walls of the container to the<br />

particle by electromagnetic radiation,with the r<strong>and</strong>om<br />

emission <strong>and</strong> absorption of the photons producing the<br />

Brownian motion. In fact,even temperature is not a<br />

particularly effective modulator of Brownian motion since<br />

the velocity of the particles depends on the square root of the<br />

temperature. So to reduce r<strong>and</strong>om thermal fluctuations to<br />

10% of the amount present at room temperature,one would<br />

have to drop the temperature from 300 K to 3 K. [12,13] It seems<br />

sensible,therefore,to try to utilize Brownian motion when<br />

designing molecular machines rather than make structures<br />

that have to fight against it. Indeed,the question of how to<br />

(<strong>and</strong> whether it is even possible to) harness the inherent<br />

r<strong>and</strong>om motion present at small length scales to generate<br />

motion <strong>and</strong> do work at larger length scales has vexed<br />

scientists for some considerable time.<br />

1.2.1. The Brownian Motion “Thought <strong>Machines</strong>”<br />

The laws of thermodynamics govern how systems gain,<br />

process,<strong>and</strong> release energy <strong>and</strong> are central to the use of<br />

particle motion to do work at any scale. The zeroth law of<br />

thermodynamics tells us about the nature of equilibrium,the<br />

first law is concerned with the total energy of a system,while<br />

the third law sets the limits against which absolute measurements<br />

can be made. Whenever energy changes h<strong>and</strong>s,<br />

however,(body to body or form to form) it is the second<br />

law of thermodynamics which comes into play. This law<br />

provides the link between the fundamentally reversible laws<br />

of physics <strong>and</strong> the clearly irreversible nature of the universe in<br />

which we exist. Furthermore,it is the second law of<br />

thermodynamics,with its often counterintuitive consequences,that<br />

governs many of the important design aspects of how<br />

to harness Brownian motion to make molecular-level<br />

machines. Indeed,the design of tiny machines capable of<br />

doing work was the subject of several celebrated historical<br />

“thought-machines” intended to test the very nature of the<br />

second law of thermodynamics. [14–18]<br />

1.2.1.1. Maxwell’s Demon<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Scottish physicist James Clerk Maxwell played a major<br />

role (along with Ludwig Boltzmann) in developing the kinetic<br />

theory of gases,which established the relationship between<br />

heat <strong>and</strong> particle motion <strong>and</strong> gave birth to the concept of<br />

statistical mechanics. In doing so,Maxwell realized the<br />

profundity of the statistical nature of the second law of<br />

thermodynamics which had recently been formulated [19] by<br />

Rudolf Clausius <strong>and</strong> William Thomson (later Lord Kelvin). In<br />

an attempt to illustrate this feature,Maxwell devised the<br />

thought experiment which has come to be known as<br />

Maxwell s demon. [14,15,20]<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

75


Reviews<br />

Maxwell envisaged a gas enclosed within a container,to<br />

<strong>and</strong> from which no heat or matter could flow. The second law<br />

of thermodynamics requires that no gradient of heat or<br />

pressure can spontaneously arise in such a system,as that<br />

would constitute a reduction in entropy. Maxwell imagined<br />

the system separated into two sections by a partition<br />

(Figure 2). Having just proven that the molecules in a gas at<br />

Figure 2. a) Maxwell’s “temperature demon” in which a gas at uniform temperature is sorted into “hot”<br />

<strong>and</strong> “cold” molecules. [15] Particles with energy higher than the average are represented by red dots while<br />

blue dots represent particles with energies lower than the average. All mechanical operations carried out<br />

by the demon involve no work—that is, the door is frictionless <strong>and</strong> it is opened <strong>and</strong> closed infinitely<br />

slowly. The depiction of the demon outside the vessel is arbitrary <strong>and</strong> was not explicitly specified by<br />

Maxwell. b) A Maxwellian “pressure demon” in which a pressure gradient is established by the door<br />

being opened only when a particle in the left compartment approaches it. [15c]<br />

a particular temperature have energies statistically distributed<br />

about a mean,Maxwell postulated a tiny “being” able to<br />

detect the velocities of individual molecules <strong>and</strong> open <strong>and</strong><br />

close a hole in the partition so as to allow molecules moving<br />

faster than the average to move in one direction (R!L in<br />

Figure 2) <strong>and</strong> molecules moving slower than the average to<br />

move in the other (L!R in Figure 2). All the time,the<br />

number of particles in each half <strong>and</strong> the total amount of<br />

energy in the system remains the same. The result of the<br />

demon s endeavors being successful would be that one end of<br />

the system (the “fast” end,L) would become hot <strong>and</strong> the<br />

other (the “slow” end,R) cold; thus a temperature gradient is<br />

set up without doing any work,contrary to the second law of<br />

thermodynamics.<br />

After its publication in “Theory of Heat”, [15b] Thomson<br />

exp<strong>and</strong>ed upon Maxwell s idea in a paper [21] read before the<br />

Royal Society of Edinburgh on 16 February 1874,<strong>and</strong><br />

published a few weeks later in Nature, [22] introducing the<br />

term “demon” for Maxwell s being. [23] In using this word,<br />

Thomson apparently did not intend to suggest a malicious<br />

imp,but rather something more in keeping with the ancient<br />

Greek roots of the word (more usually daemon) as a<br />

supernatural being of a nature between gods <strong>and</strong> humans.<br />

Indeed,neither Maxwell nor Thomson saw the demon as a<br />

threat to the second law of thermodynamics,but rather an<br />

illustration of its limitations—an exposition of its statistical<br />

nature. This,however,has not been the view of many<br />

subsequent investigators,who have perceived the demon as<br />

an attempt to construct a perpetual motion machine driven by<br />

thermal fluctuations. The term “Maxwell s demon” has come<br />

to describe all manner of hypothetical constructs designed to<br />

overcome the second law of thermodynamics by continually<br />

extracting energy to do work from the thermal bath. [14]<br />

Maxwell noted that the demon principle could be<br />

demonstrated in a number of different ways; a “pressure”<br />

demon,for example,(Figure 2b) could sort particles so that<br />

more end up in one end of a vessel than the other,which<br />

required different information to<br />

operate from the original temperature<br />

demon (the direction of<br />

approach of a particle,not its<br />

speed).<br />

1.2.1.2. Szilard’s Engine<br />

Both Maxwell <strong>and</strong> Thomson<br />

appreciated that the operation of<br />

these systems for separating<br />

Brownian particles appeared to<br />

rely on the demon s “intelligence”<br />

as an animate being,but<br />

did not try to quantify it. Leo<br />

Szilard made the first attempt to<br />

mathematically relate the<br />

demon s intelligence to the thermodynamics<br />

of the process by<br />

considering the performance of a<br />

machine based on the “pressure<br />

demon”,namely,Szilard s engine<br />

(Figure 3). [17] Szilard realized that the operations which the<br />

demon carries out can be reduced to a simple computational<br />

process. In particular,he recognized the process requires<br />

measurement of the approach of the particle to gain<br />

information about its direction <strong>and</strong> speed,which must be<br />

remembered <strong>and</strong> then acted upon. [24]<br />

1.2.1.3. Smoluchowski’s Trapdoor<br />

The concept of a purely mechanical Brownian motion<br />

machine which does not require any intelligent being to<br />

operate it was first explored by Marian von Smoluchowski<br />

who imagined the Maxwell system as two gas-containing<br />

compartments with a spring-loaded trapdoor in place of the<br />

demon-operated hole (Figure 4a). [16] If the spring is weak<br />

enough,it might appear that the door could be opened by<br />

collisions with gas molecules moving in one direction (L!R<br />

in Figure 4a) but not the other,<strong>and</strong> so allow transport of<br />

molecules preferentially in one direction thereby creating a<br />

pressure gradient between the two compartments. Smoluchowski<br />

recognized (although could not prove) that if the<br />

door had no way of dissipating the energy it gains from<br />

Brownian collisions it would be subject to the same amount of<br />

r<strong>and</strong>om thermal motion as the rest of the system <strong>and</strong> would<br />

not then function as the desired one-way valve.<br />

1.2.1.4. Feynman’s Ratchet <strong>and</strong> Pawl<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Richard Feynman revisited these ideas in his celebrated<br />

discussion of a miniature ratchet <strong>and</strong> pawl—a construct<br />

76 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 3. Szilard’s engine which utilizes a “pressure demon”. [17]<br />

a) Initially a single Brownian particle occupies a cylinder with a piston<br />

at either end. A frictionless partition is put in place to divide the<br />

container into two compartments (a!b). b) The demon then detects<br />

the particle <strong>and</strong> determines in which compartment it resides. c) Using<br />

this information, the demon is able to move the opposite piston into<br />

position without meeting any resistance from the particle. d) The<br />

partition is removed, allowing the “gas” to exp<strong>and</strong> against the piston,<br />

doing work against any attached load (e). To replenish the energy used<br />

by the piston <strong>and</strong> maintain a constant temperature, heat must flow<br />

into the system. To complete the thermodynamic cycle <strong>and</strong> reset the<br />

machine, the demon’s memory of where the particle was must be<br />

erased (f!a). To fully justify the application of a thermodynamic<br />

concept such as entropy to a single-particle model, a population of<br />

Szilard devices is required. The average for the ensemble over each of<br />

these devices can then be considered to represent the state of the<br />

system, which is comparable to the time average of a single multiparticle<br />

system at equilibrium, in a fashion similar to the statistical<br />

mechanics derivation of thermodynamic quantities.<br />

designed to illustrate how the irreversible second law of<br />

thermodynamics arises from the fundamentally reversible<br />

laws of motion. [18] Feynman s device (Figure 4b) consists of a<br />

miniature axle with vanes attached to one end,surrounded by<br />

a gas at temperature T 1. At the other end of the axle is a<br />

ratchet <strong>and</strong> pawl system,held at temperature T 2. The question<br />

posed by the system is whether the r<strong>and</strong>om oscillations<br />

produced by gas molecules bombarding the vanes can be<br />

rectified by the ratchet <strong>and</strong> pawl so as to get net motion in one<br />

direction. Exactly analogous to Smoluchowski s trapdoor,<br />

Feynman showed that if T 1 = T 2 then the ratchet <strong>and</strong> pawl<br />

cannot extract energy from the thermal bath to do work.<br />

Feynman s major contribution from the perspective of<br />

molecular machines,however,was to take the analysis one<br />

stage further: if such a system cannot use thermal energy to do<br />

work,what is required for it to do so? Feynman showed that<br />

when the ratchet <strong>and</strong> pawl are cooler than the vanes (that is,<br />

T 1 > T 2) the system does indeed rectify thermal motions <strong>and</strong><br />

can do work (Feynman suggested lifting a hypothetical flea<br />

attached by a thread to the ratchet). [25] Of course,the machine<br />

now does not threaten the second law of thermodynamics as<br />

dissipation of heat into the gas reservoir of the ratchet <strong>and</strong><br />

pawl occurs,so the temperature difference must be maintained<br />

by some external means. Although insulating a<br />

molecular-sized system from the outside environment is<br />

Figure 4. a) Smoluchowski’s trapdoor: an “automatic” pressure<br />

demon (the directionally discriminating behavior is carried out by a<br />

wholly mechanical device, a trapdoor which is intended to open when<br />

hit from one direction but not the other). [16] Like the pressure demon<br />

shown in Figure 2b, Smoluchowski’s trapdoor aims to transport<br />

particles selectively from the left compartment to the right. However,<br />

in the absence of a mechanism whereby the trapdoor can dissipate<br />

energy it will be at thermal equilibrium with its surroundings. This<br />

means it must spend much of its time open, unable to influence the<br />

transport of particles. Rarely, it will be closed when a particle<br />

approaches from the right <strong>and</strong> will open on collision with a particle<br />

coming from the left, thus doing its job as intended. Such events are<br />

balanced, however, by the door snapping shut on a particle from the<br />

right, pushing it into the left chamber. Overall, the probability of a<br />

particle moving from left to right is equal to that for moving right to<br />

left <strong>and</strong> so the trapdoor cannot accomplish its intended function<br />

adiabatically. b) Feynman’s ratchet <strong>and</strong> pawl. [18] It might appear that<br />

Brownian motion of the gas molecules on the paddle wheel in the<br />

right-h<strong>and</strong> compartment can do work by exploiting the asymmetry of<br />

the teeth on the cog of the ratchet in the left-h<strong>and</strong> compartment.<br />

While the spring holds the pawl between the teeth of the cog, it does<br />

indeed turn selectively in the desired direction. However, when the<br />

pawl is disengaged, the cog wheel need only r<strong>and</strong>omly rotate a tiny<br />

amount in the other direction to move back one tooth whereas it must<br />

rotate r<strong>and</strong>omly a long way to move to the next tooth forward. If the<br />

paddle wheel <strong>and</strong> ratchet are at the same temperature (that is, T 1 = T 2)<br />

these rates cancel out. However, if T 1 ¼6 T 2 then the system will<br />

directionally rotate, driven solely by the Brownian motion of the gas<br />

molecules. Part (b) reprinted with permission from Ref. [18].<br />

difficult (<strong>and</strong> temperature gradients cannot be maintained<br />

over molecular-scale distances,see Section 1.3),what this<br />

hypothetical construct provides is the first example of a<br />

plausible mechanism for a molecular motor—whereby the<br />

r<strong>and</strong>om thermal fluctuations characteristic of this size regime<br />

are not fought against but instead are harnessed <strong>and</strong> rectified.<br />

The key ingredient is the external addition of energy to the<br />

system,not to generate motion but rather to continually or<br />

cyclically drive the system away from equilibrium,thereby<br />

maintaining a thermally activated relaxation process that<br />

directionally biases Brownian motion towards equilibrium. [26]<br />

This profound idea is the key to the design of molecular-level<br />

systems that work through controling Brownian motion <strong>and</strong> is<br />

exp<strong>and</strong>ed upon in Section 1.4.<br />

1.2.2. <strong>Machines</strong> That Operate at Low Reynolds Number<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Whilst rectifying Brownian motion may provide the key to<br />

powering molecular-level machines,it tells us nothing about<br />

how that power can be used to perform tasks at the nanoscale<br />

<strong>and</strong> what tiny mechanical machines can <strong>and</strong> cannot be<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

77


Reviews<br />

expected to do. The constant presence of Brownian motion is<br />

not the only distinction between motion at the molecular level<br />

<strong>and</strong> in the macroscopic world. In the macroscopic world,the<br />

equations of motion are governed by inertial terms (dependent<br />

on mass). Viscous forces (dependent on particle<br />

dimensions) dampen motion by converting kinetic energy<br />

into heat,<strong>and</strong> objects do not move until they are provided<br />

with specific energy to do so. In a macroscopic machine,this is<br />

often provided through a directional force when work is done<br />

to move mechanical components in a particular way. As<br />

objects become less massive <strong>and</strong> smaller in dimensions,<br />

inertial terms decrease in importance <strong>and</strong> viscosity begins to<br />

dominate. A parameter which quantifies this effect is the<br />

Reynolds number R,which is essentially the ratio of inertial<br />

to viscous forces,<strong>and</strong> is given by Equation (1) for a particle of<br />

length a that is moving at velocity v in a medium with viscosity<br />

h <strong>and</strong> density 1. [27]<br />

R ¼ av1<br />

h<br />

Size affects the modes of motion long before the nanoscale<br />

is reached. Even at the mesoscopic level of bacteria<br />

(length dimensions ca. 10 5 m),viscous forces dominate. At<br />

the molecular level,the Reynolds number is extremely low<br />

(except at low pressures in the gas phase or,possibly,in the<br />

free volume within rigid frameworks in the solid state) <strong>and</strong><br />

the result is that molecules,or their components,cannot be<br />

given a one-off “push” in the macroscopic sense. Thus,<br />

momentum is irrelevant. The motion of a molecular-level<br />

object is determined entirely by the forces acting on it at that<br />

particular instant—whether they are externally applied<br />

forces,viscosity,or r<strong>and</strong>om thermal perturbations <strong>and</strong><br />

Brownian motion. Furthermore,the tiny masses of nanoscopic<br />

objects means that the force of gravity is insignificant<br />

for them. Since the physics which governs mechanical<br />

dynamic processes in the two size regimes is completely<br />

different,macroscopic <strong>and</strong> nanoscale motors require fundamentally<br />

different mechanisms for controlled transport or<br />

propulsion. Moreover,the high ratios of surface area to<br />

volume for molecules mean they are inherently sticky <strong>and</strong> this<br />

will have a profound effect on how molecular-sized machines<br />

are organized <strong>and</strong> interact with one another. In general terms,<br />

this analysis leads to a central tenet: while the macroscopic<br />

machines we encounter in everyday life may provide the<br />

inspiration for what we might like molecular machines to<br />

achieve,drawing too close an analogy for how they might do it<br />

may not be the best design strategy. The “rules of the game”<br />

at large <strong>and</strong> small length scales are simply too different.<br />

[1m,12,13,28]<br />

1.3. Lessons to Learn from Biology<br />

Help is at h<strong>and</strong>,however,because despite all these<br />

problems,we know that motors <strong>and</strong> machines at the<br />

molecular level are conceptually feasible—they are already<br />

all around us. Nature has developed a working molecular<br />

nanotechnology that it employs to astonishing effect in<br />

ð1Þ<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

virtually every significant biological process. [11] Appreciating<br />

in general terms how nature has overcome the issues of scale,<br />

environment,equilibrium,Brownian motion,<strong>and</strong> viscosity is<br />

useful for indicating general traits for the design of synthetic<br />

molecular machine systems <strong>and</strong> how they might be used.<br />

The membranes which encase cells <strong>and</strong> their organelles<br />

allow the compartmentalization of cellular processes <strong>and</strong><br />

constituents,thereby maintaining the non-equilibrium distributions<br />

essential for life. These lipophilic barriers contain a<br />

diverse range of functionality which facilitate the movement<br />

of various ionic <strong>and</strong> polar species through channels,through<br />

relays or the use of mobile carrier species. [29] A number of<br />

different mechanisms are employed to power motion from<br />

one compartment to another. Besides passive diffusion down<br />

a concentration gradient in the transported species,an<br />

electrochemical gradient (a transmembrane electrical potential,a<br />

concentration gradient,or,most commonly,a combination<br />

of the two) originating from another species can be<br />

used. Such “gradient pumping” mechanisms move one<br />

species against its concentration gradient at the expense of<br />

another gradient. The motion of the two species can be in the<br />

same direction (symport) or in opposite directions (antiport).<br />

If the sole power source is an electrical potential,only one<br />

species crosses the membrane (a uniport mechanism). These<br />

processes are governed by sophisticated control mechanisms<br />

which open or close highly selective channels,carriers,<strong>and</strong> cotransporters<br />

in response to different stimuli,yet they all<br />

operate by relaxation down a directional transmembrane<br />

electrochemical gradient towards the thermodynamic equilibrium<br />

position.<br />

The primary concentration gradients which are used to<br />

power subsequent secondary transport processes are maintained<br />

by a smaller number of ion pumps. These transmembrane<br />

mechanical devices convert nondirectional chemical<br />

reactions (the most abundant family,the ATPases,<br />

harness the hydrolysis of ATP) into vectorial transport of<br />

ionic species against an electrochemical gradient. The precise<br />

mechanisms by which these devices operate are still a subject<br />

of intense investigation. However,certain principles are<br />

becoming apparent. In particular,changes in the binding<br />

affinity at selective sites in the transmembrane region of the<br />

pump must be coupled to conformational changes which also<br />

modulate access to these sites from either side of the<br />

membrane so that only motion in the desired direction is<br />

achieved. [30]<br />

A recent series of structures has outlined how just such a<br />

mechanism operates in the sarcoplasmic reticulum Ca 2+ -<br />

ATPase pump. [31] Calcium ions bind to two high affinity sites<br />

which are open to the cytoplasm (B,Figure 5). Phosphorylation<br />

of the enzyme by ATP then results in a conformational<br />

change which shuts off access from either side (ratcheting [32] ).<br />

Release of ADP is concomitant with a further conformational<br />

change which opens up access to the lumen side <strong>and</strong> disrupts<br />

the Ca 2+ -binding residues (A,Figure 5). The calcium ions are<br />

released (escapement [32] ) <strong>and</strong> swapped for protons which are<br />

subsequently occluded from the lumen by binding of a water<br />

molecule. Release of phosphate occurs with regeneration of<br />

the starting state,once again making the high-affinity Ca 2+ -<br />

binding sites accessible to the cytoplasm. Similar mechanisms<br />

78 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 5. Crystal structures of the calcium-binding region in<br />

Ca 2+ -ATPase in both low Ca 2+ -affinity (A) <strong>and</strong> high Ca 2+ -affinity (B)<br />

conformations. [31] Calcium ions are shown in green, <strong>and</strong> the numbers<br />

correspond to specific protein helices. The crystal structure pictures<br />

are reprinted with permission from Ref. [30b].<br />

are thought to be responsible for the operation of the other<br />

members in the ATPase family.<br />

The pumping of protons across the energy-transducing<br />

membranes of mitochondria,chloroplasts,<strong>and</strong> photosynthetic<br />

<strong>and</strong> respiratory bacteria is a particularly important process as<br />

it generates the protonmotive force necessary to operate ATP<br />

synthase (<strong>and</strong> also directly or indirectly powers the secondary<br />

transport of other species across these membranes). The ways<br />

in which nature achieves this pumping function are surprisingly<br />

diverse (by contrast,ATP synthase is highly conserved<br />

across organisms). [33] In most examples,charge-separated<br />

states <strong>and</strong> electron-transfer processes are used to generate<br />

transmembrane electric potentials. Overall,therefore,a<br />

directional electrochemical driving force is set up to drive<br />

the translocation of protons,somewhat similar to the gradient<br />

pumping secondary transport processes discussed above.<br />

Proton transport is concomitant with the resetting of the<br />

redox centers in processes often termed “redox loops” or<br />

“Mitchell loops”. [34] Only in the case of bacteriorhodopsin<br />

(the photosynthetic center of certain purple bacteria) does it<br />

appear that light energy is directly converted into protonmotive<br />

force by a purely conformational pumping mechanism. [35]<br />

Again,the details of this process are not yet fully understood,<br />

but the correlation of conformational changes that control the<br />

direction in which the protons can move with changes in the<br />

affinity (that is,pK a) of binding sites are likely to be<br />

involved. [36]<br />

Nature also uses molecular machines to transport objects<br />

around within cells (for example,kinesin or dynein); to move<br />

whole organisms or their parts (for example,myosin or the<br />

bacterial flagellar motor); to process DNA <strong>and</strong> RNA (for<br />

example,helicases or polymerases); to convert protonmotive<br />

force into the synthesis of ATP (ATP synthase); <strong>and</strong> for may<br />

other functions besides. In these cases,too,much has been<br />

learnt about the various ways in which they perform their<br />

remarkable tasks,yet much remains to be discovered about<br />

each before the precise mechanisms can be elucidated. [11]<br />

Accordingly,we can see that there are many general <strong>and</strong><br />

fundamental differences between biological molecular<br />

machines <strong>and</strong> the man-made machines of the macroscopic<br />

world. Listed below are what appear to us to be the most<br />

Angew<strong>and</strong>te<br />

Chemie<br />

significant aspects of biological machines to bear in mind<br />

when considering synthetic molecular machine design.<br />

1) Biological machines are soft,not rigid.<br />

2) Biological systems operate at near-ambient temperatures<br />

(heat is dissipated almost instantaneously at small length<br />

scales so they cannot exploit temperature gradients).<br />

3) Biological motors utilize chemical energy,in the form of<br />

covalent-bond breaking/formation of high-energy compounds<br />

such as ATP,NADH,<strong>and</strong> NADPH or concentration<br />

gradients.<br />

4) Biomachines operate in solution,or at surfaces,under<br />

conditions of intrinsically high viscosity.<br />

5) Nature utilizes—rather than opposes—Brownian<br />

motion. Biomolecular machines need not use chemical<br />

energy to initiate movement—their components are<br />

constantly in motion—rather,they function by manipulating<br />

(ratcheting) that movement. Furthermore,constant<br />

thermal motion <strong>and</strong> small “reaction vessels” (cells<br />

<strong>and</strong> their organelles) ensure that the mixing of biological<br />

machines,their substrates,<strong>and</strong> their fuel is extremely<br />

rapid,in spite of the high viscosity they experience.<br />

6) The viscous environment <strong>and</strong> constant thermal motion<br />

mean that biological machines have no use for smooth,<br />

low-friction surfaces. Genuinely smooth features are not<br />

possible on the molecular scale,of course,since the<br />

machine-component dimensions are close to the dimensions<br />

of the intrinsic unit of matter—the atom.<br />

7) Biomotors <strong>and</strong> other mobile machines utilize architectures<br />

(for example,tracks) which serve to restrict most of<br />

the degrees of freedom of the machine components <strong>and</strong>/<br />

or the substrate(s) they act upon. The molecular machine<br />

<strong>and</strong> the substrate(s) it is acting upon remain kinetically<br />

associated during the operation of the machine. For<br />

example,kinesin only functions as it “walks” processively<br />

down a microtubule,not by simply binding to the<br />

microtubule at different places (that is,not by completely<br />

detaching,exchanging with the bulk,<strong>and</strong> rebinding).<br />

Similarly,pumps work by internally compartmentalizing<br />

ions so that they cannot exchange with others outwith the<br />

machine.<br />

8) The operation <strong>and</strong> structure of biological machines are<br />

governed by noncovalent interactions (intramolecular<br />

<strong>and</strong> intermolecular),many of which exploit the aqueous<br />

environment in which the machines assemble <strong>and</strong><br />

operate.<br />

9) Biological machines are made by combinations of multiple<br />

parallel synthesis <strong>and</strong> self-assembly processes utilizing<br />

a relatively small range of building blocks: amino<br />

acids nucleic acids,lipids,<strong>and</strong> saccharides.<br />

10) Living organisms intrinsically <strong>and</strong> necessarily function<br />

far from equilibrium—a state maintained by the compartmentalization<br />

of processes into cells,vesicles,<strong>and</strong><br />

organelles.<br />

We should also remember,however,that the mechanisms<br />

<strong>and</strong> function of biomolecular machines are restricted by<br />

evolutionary constraints. Over billions of years,evolution has<br />

led to machines of a complexity not yet possible through<br />

rational design,but it intrinsically proceeds through small<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

79


Reviews<br />

iterations <strong>and</strong> tends to retain the first successful solution<br />

arrived at for each problem. The human molecular-machine<br />

designer has at his or her disposal a much larger chemical<br />

toolbox,a wider range of possible operating conditions,<strong>and</strong>,<br />

as we shall see,a design approach that favors innovation.<br />

Furthermore,whilst an appreciation of the characteristics<br />

of biological machines can give us broad clues about how to<br />

make molecular-level devices that exploit mechanical motion,<br />

we underst<strong>and</strong> few of the precise details of how their designs<br />

work. How does an inherently mechanically directionless<br />

chemical reaction—the conversion of ATP into ADP—drive<br />

the directional motion of calcium ions by Ca 2+ -ATPase?<br />

What is the detailed nature of the conformational changes<br />

which can set-up or disrupt the highly selective binding sites?<br />

Why exactly do the movements of the individual peptide<br />

subunits that cause this to happen occur in that particular<br />

sequence? What are the details of the pathways connecting<br />

the binding sites with the outside? What is the role of the<br />

kinetics <strong>and</strong> thermodynamics of each amino acid in this<br />

mechanism? In fact,biological motors <strong>and</strong> machines are so<br />

complex that studying simple synthetic chemical systems may<br />

help in the underst<strong>and</strong>ing of exactly how they work. Likewise,<br />

fully underst<strong>and</strong>ing the biological systems can only aid the<br />

design of sophisticated artificial systems. At present,however,<br />

much of the most detailed information for devising the basic<br />

mechanisms for synthetic molecular machines comes not from<br />

biology,but from non-equilibrium statistical mechanics.<br />

1.4. Lessons To Learn from Physics, Mathematics, <strong>and</strong> Statistical<br />

Mechanics<br />

1.4.1. Breaking Detailed Balance<br />

The Principle of detailed balance [37] states that at equilibrium,transitions<br />

between any two states take place in either<br />

direction at the same rate so that no net flux is generated. This<br />

rules out the maintenance of equilibria by cyclic processes<br />

such as A!B!C!A rather than AÐB + BÐC + CÐA [12]<br />

<strong>and</strong> is a formal indication that machines such as Smoluchowski<br />

s trapdoor <strong>and</strong> Feynman s ratchet <strong>and</strong> pawl cannot<br />

operate adiabatically. However,in an out-of-equilibrium<br />

system (for example,if T 1 ¼6 T 2 for the Feynman ratchet <strong>and</strong><br />

pawl),“detailed balance” is broken,<strong>and</strong> net work can be done<br />

by the fluxional exchange process as the system moves<br />

spontaneously towards equilibrium.<br />

To underst<strong>and</strong> the mechanisms by which mechanical<br />

machines can operate on length scales at which Brownian<br />

motion occurs it is useful to appreciate precisely why work<br />

can be done by a r<strong>and</strong>om exchange process between two<br />

states only while detailed balance is broken. A simple analogy<br />

helps illustrate why detailed balance holds in a fluxionally<br />

exchanging system at equilibrium <strong>and</strong> the requirements<br />

necessary to break it. Consider a pair of scales with many<br />

ants running r<strong>and</strong>omly between the two sides,the scales will<br />

be statistically balanced (but not necessarily horizontal,that<br />

only occurs if the pivot of the scale is positioned in the middle)<br />

<strong>and</strong> remain more or less steady at a stable angle. Even if a<br />

barrier is suddenly placed between the scales,thus stopping<br />

the two sides being in equilibrium,they remain balanced.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

However,if ants are added to one side or the other the scales<br />

become unbalanced <strong>and</strong> will tip accordingly. If the barrier is<br />

removed,the ants (assuming they can overcome the effect of<br />

gravity!) resume scurrying between the two sides <strong>and</strong><br />

eventually equilibrium—<strong>and</strong> balance—will be restored.<br />

Note that even if we start without the barrier in place,if<br />

ants are suddenly added to one side balance will be broken for<br />

some time—<strong>and</strong> the scale will tip—until the roaming ants<br />

establish equilibrium <strong>and</strong> balance is restored. The breaking of<br />

balance allows a task to be performed as the exchanging<br />

system returns toward equilibrium. The net displacement that<br />

arises as the scales tip over (or as they right themselves) could<br />

be used to lift a feather to which the scale was attached.<br />

Breaking detailed balance is the key to doing work with a<br />

machine that operates at length scales on which Brownian<br />

motion occurs.<br />

During the past decade,a number of remarkable theoretical<br />

formalisms have been developed in mathematics<br />

(particularly game theory) <strong>and</strong> non-equilibrium statistical<br />

physics that explain how the directional transport of Brownian<br />

particles over periodic potential-energy surfaces can<br />

occur from unbiased periodic or stochastic perturbations of<br />

the system. [38,39] Underlying each of these Brownian ratchet or<br />

motor mechanisms are three components: [39b] 1) a r<strong>and</strong>omizing<br />

element; [40] 2) an energy input to avoid falling foul of the<br />

second law of thermodynamics; <strong>and</strong> 3) asymmetry in the<br />

energy or information potential in the dimension in which the<br />

motion occurs. [41] It is now widely accepted that ratchet<br />

mechanisms have a central role in explaining the operation of<br />

biological motor proteins. [42,43] They have also been successfully<br />

used to develop transport <strong>and</strong> separation devices for<br />

mesoscopic particles <strong>and</strong> macromolecules,microfluidic<br />

pumping,as well as quantum <strong>and</strong> electronic applications. [43,44]<br />

For molecular-level motors <strong>and</strong> other machines,the necessary<br />

r<strong>and</strong>omizing element can be Brownian motion of the<br />

substrate or the machine components. The other two requirements<br />

(energy <strong>and</strong> anisotropy) can be supplied in different<br />

ways according to the type of fluctuation-driven transport<br />

mechanism.<br />

1.4.2. Fluctuation-Driven Transport Mechanisms<br />

Many different possible types of theoretical fluctuationdriven<br />

(“Brownian ratchet”) mechanisms have been proposed,including<br />

flashing ratchets,tilting ratchets,Seebeck<br />

ratchets,drift ratchets,Hamiltonian ratchets,temperature<br />

ratchets,<strong>and</strong> entropic ratchets. [39] They can be classified <strong>and</strong><br />

grouped together in many different (not always mutually<br />

exclusive) ways,<strong>and</strong> mechanisms that might be indistinguishable<br />

in chemical terms can be differentiated because of the<br />

nature of the physics involved. Whilst it is not clear (at least to<br />

us!) how some of the theoretical mechanisms could be<br />

applied to molecular structures,many offer clear design<br />

opportunities for the synthetic chemist. The details of the<br />

various mechanisms have been discussed extensively in the<br />

physics literature,so we will limit our discussion here to some<br />

specific variations which appear particularly well-suited for<br />

chemical systems. We choose to distinguish between two<br />

overarching classes of mechanism: energy ratchets,which fall<br />

80 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

into two basic types—pulsating ratchets <strong>and</strong> tilting ratchets—<br />

<strong>and</strong> are the subject of a recent major review by Reimann, [39f]<br />

<strong>and</strong> information ratchets,which are much less common in the<br />

physics literature,but have been discussed by Parrondo,<br />

Astumian,<strong>and</strong> others. [39h,42b,45] Both energy ratchets <strong>and</strong><br />

information ratchets bias the movement of a Brownian<br />

substrate. [46] However,we will also show (Sections 1.4.3 <strong>and</strong><br />

4.4) that they offer clues for how to go beyond a simple switch<br />

with a chemical machine to enable tasks to be performed<br />

through the non-equilibrium control of conformational <strong>and</strong><br />

co-conformational changes within molecular structures.<br />

[39f ]<br />

1.4.2.1. Pulsating Ratchets<br />

Pulsating ratchets are a general category of energy ratchet<br />

in which potential-energy minima <strong>and</strong> maxima are varied in a<br />

periodic or stochastic fashion,independent of the position of<br />

the particle on the potential-energy surface. In its simplest<br />

form this can be considered as an asymmetric sawtooth<br />

potential being repetitively turned on <strong>and</strong> off faster than<br />

Brownian particles can diffuse over more than a small fraction<br />

of the potential energy surface (an “on–off” ratchet,<br />

Figure 6). The result is net directional transport of the<br />

particles across the surface (left to right in Figure 6).<br />

More general than the special case of an on–off ratchet,<br />

any asymmetric periodic potential may be regularly or<br />

stochastically varied to give a ratchet effect (such mechanisms<br />

are generally termed “fluctuating potential” ratchets). As<br />

with the simple on–off ratchet,most commonly encountered<br />

examples involve switching between two different potentials<br />

<strong>and</strong> are therefore often termed “flashing” ratchets. A classic<br />

example,which has particular relevance for explaining a<br />

Figure 6. An example of a pulsating ratchet mechanism—an on–off<br />

ratchet. [39f] a) The Brownian particles start out in energy minima on the<br />

potential-energy surface with the energy barriers @ k BT. b) The potential<br />

is turned off so that free Brownian motion powered diffusion is<br />

allowed to occur for a short time period (much less than required to<br />

reach global equilibrium). c) On turning the potential back on again,<br />

the asymmetry of the potential means that the particles have a greater<br />

probability of being trapped in the adjacent well to the right rather<br />

than the adjacent well to the left. Note this step involves raising the<br />

energy of the particles. d) Relaxation to the local energy minima<br />

(during which heat is emitted) leads to the average position of the<br />

particles moving to the right. Repeating steps (b)–(d) progressively<br />

moves the Brownian particles further <strong>and</strong> further to the right.<br />

number of biological processes [42b] as well as being the basis<br />

for a [2]catenane rotary motor (see Section 4.6.3),is illustrated<br />

in Figure 7. It consists in physical terms of an<br />

asymmetric potential-energy surface (comprising a periodic<br />

Figure 7. Another example of a pulsating ratchet mechanism—a flashing<br />

ratchet. [42b] For details of its operation, see the text.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

series of two different minima <strong>and</strong> two different maxima)<br />

along which a Brownian particle is directionally transported<br />

by sequentially raising <strong>and</strong> lowering each set of minima <strong>and</strong><br />

maxima. The particle starts in a green or orange well<br />

(Figure 7a or c). Raising that energy minimum while lowering<br />

those in adjacent wells provides the impetus for the particle to<br />

change position by Brownian motion (Figure 7b!7cor7d!<br />

7e). By simultaneously (or beforeh<strong>and</strong>) changing the relative<br />

heights of the energy barrier to the next energy well,the<br />

kinetics of the Brownian motion in each direction are<br />

different <strong>and</strong> the particle is transported from left to right.<br />

Note that the position of the particle does not influence the<br />

sequence in which (or when,or if) the energy minima <strong>and</strong><br />

maxima are changed. Furthermore,the switching of the<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

81


Reviews<br />

potential does not have to be regular. As long as the two sets<br />

of energy minima <strong>and</strong> maxima are repetitively switched in the<br />

same order,the particle will tend to be transported from left<br />

to right in Figure 7,even though occasionally it may move<br />

over a barrier in the wrong direction.<br />

The basic pulsating ratchet requirements can also be<br />

realized in another way. A potential such as that shown in<br />

Figure 7a can be given a directional drift velocity. Such<br />

systems are often termed “traveling potential” ratchets. This<br />

principle is essentially the same as macroscopic devices such<br />

as Archimedes screw,yet in the presence of thermal<br />

fluctuations these systems exhibit Brownian ratchet characteristics<br />

<strong>and</strong> are closely related to tilting the potential in one<br />

direction (as discussed in Section 1.4.2.2). Clearly,however,<br />

an asymmetric potential is not absolutely required,nor in fact<br />

are thermal fluctuations; imagine,for example,a particle<br />

“surfing” on a traveling wave on the surface of a liquid. This<br />

category of mechanism is at the boundary between fluctuation-driven<br />

transport <strong>and</strong> transport as a result of potential<br />

gradients,with the precise location on this continuum<br />

dependent on the importance of thermal fluctuations <strong>and</strong><br />

spatial asymmetry under the conditions chosen. In terms of<br />

chemical systems,the traveling-potential mechanism has most<br />

relevance for the field-driven processes discussed in Section 5<br />

<strong>and</strong> the self-propulsion mechanisms of Section 6,while in<br />

Section 7 we shall discuss some situations in which the<br />

balance between r<strong>and</strong>om fluctuations <strong>and</strong> external directional<br />

driving force is shifted strongly towards the latter. The<br />

traveling motion does not have to be continuous,but rather<br />

may take place in discrete steps. Furthermore,arranging a<br />

continuously traveling potential to “jump” distances which<br />

are multiples of its period,at r<strong>and</strong>om or regular intervals,can<br />

be used to escape from the inherent directionality of the<br />

traveling-wave scheme <strong>and</strong> it has been shown that the ratchet<br />

dynamics are essentially unaffected. In the limiting situation,<br />

this can be reduced to dichotomous switching between two<br />

spatially shifted potentials which are otherwise identical,<br />

which is very similar to a rocking ratchet (see Section 1.4.2.2)<br />

<strong>and</strong> is also relevant to the unidirectional motors discussed in<br />

Section 2.2.<br />

[39f ]<br />

1.4.2.2. Tilting Ratchets<br />

In this category,the underlying intrinsic potential remains<br />

unchanged <strong>and</strong> the detailed balance is broken by application<br />

of an unbiased driving force to the Brownian particle. Perhaps<br />

the most apparent unbiased driving force is heat,<strong>and</strong> ratchet<br />

mechanisms based on periodic or stochastic temperature<br />

variations are generally termed “temperature” or “diffusion”<br />

ratchets. In its simplest form (Figure 8) this mechanism is very<br />

similar to the on–off ratchet. Initially the thermal energy is<br />

low so that the particles cannot readily cross the energy<br />

barriers. A sudden increase in temperature can be applied so<br />

that k BT is much greater that the potential amplitude causing<br />

the particles to diffuse as if over a virtually flat potentialenergy<br />

surface (Figure 8b). Returning to the original,lower,<br />

temperature (Figure 8c) is equivalent to turning the potential<br />

back on in Figure 6c <strong>and</strong> more particles will have moved to<br />

the next well to the right than to the left. Under this scheme,<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 8. A temperature (or diffusion) ratchet. [39f] a) The Brownian<br />

particles start out in energy minima on the potential-energy surface<br />

with the energy barriers @ k BT 1. b) The temperature is increased so<br />

that the height of the barriers is ! k BT 2 <strong>and</strong> effectively free diffusion is<br />

allowed to occur for a short time period (much less than required to<br />

reach global equilibrium). c) The temperature is lowered to T 1 once<br />

more, <strong>and</strong> the asymmetry of the potential means that each particle is<br />

statistically more likely to be captured by the adjacent well to the right<br />

rather than the well to the left. d) Relaxation to the local energy<br />

minimum (during which heat is emitted) leads to the average position<br />

of the particles moving to the right. Repeating steps (b)–(d), progressively<br />

moves the Brownian particles further <strong>and</strong> further to the right.<br />

Note the similarities between this mechanism <strong>and</strong> that of the on–off<br />

ratchet shown in Figure 6.<br />

therefore,it seems that applying temperature variations to<br />

any process which involves an asymmetric potential-energy<br />

surface could result in a ratchet effect. [47] In chemical terms,<br />

this means that the rotation of a chiral group around a<br />

covalent bond can,at least in principle,be made unidirectional<br />

in such a manner. This concept is explored further in<br />

relation to a molecular ratchet system in Section 2.1.2.<br />

An unbiased driving force can also be achieved by<br />

applying a directional force in a periodic manner so that,<br />

over time,the bias averages to zero,thus generating a<br />

“rocking” ratchet. The simplest form of this is shown in<br />

Figure 9. Periodic application (to the left <strong>and</strong> then right) of a<br />

driving force that allows the particle to surmount the barriers<br />

(for example,by applying an external field if the particle is<br />

charged) results in transport. Motion over the steep barrier is<br />

again most likely as it involves the shortest distance. Such a<br />

mechanism is physically equivalent to tilting the ratchet<br />

potential in one direction then the other. Of course,if the<br />

driving force is strong enough <strong>and</strong> is constantly applied in one<br />

direction or the other,thermal fluctuations are not necessary,<br />

which would then correspond to a power stroke.<br />

Analogues of a rocking ratchet in which the applied<br />

driving force is a form of stochastic noise are known as<br />

“fluctuating force” ratchets (in certain cases,also “correlation”<br />

ratchets). Finally,a tilting ratchet can be achieved in a<br />

symmetric potential if the perturbation itself results in spatial<br />

asymmetry (becoming very similar to the travelling-potential<br />

models in Section 1.4.2.1). This may be rather clear for the<br />

82 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 9. A rocking ratchet. [39f] a) The Brownian particles start out in<br />

energy minima on the potential-energy surface with the energy barriers<br />

@ k BT. b) A directional force is applied to the left. c) An equal <strong>and</strong><br />

opposite directional force is applied to the right. d) Removal of the<br />

force <strong>and</strong> relaxation to the local energy minimum leads to the average<br />

position of the particles moving to the right. Repeating steps (b)–(d)<br />

progressively moves the Brownian particles further <strong>and</strong> further to the<br />

right.<br />

periodic force cases (imagine applying the electric field<br />

discussed above for longer in one direction than the other),<br />

but is less so for stochastic driving forces. In general,these<br />

mechanisms are known as “asymmetrically tilting” ratchets.<br />

[39h,42b, 45]<br />

1.4.2.3. Information Ratchets<br />

In the pulsating <strong>and</strong> tilting types of energy ratchet<br />

mechanisms,perturbations of the potential-energy surface—<br />

or of the particle s interaction with it—are applied globally<br />

<strong>and</strong> independent of the particle s position,while the periodicity<br />

of the potential is unchanged. Information ratchets<br />

(Figure 10) transport a Brownian particle by changing the<br />

effective kinetic barriers to Brownian motion depending on<br />

the position of the particle on the surface. In other words,the<br />

heights of the maxima on the potential-energy surface change<br />

according to the location of the particle (this requires<br />

information to be transferred from the particle to the surface)<br />

whereas the potential-energy minima do not necessarily need<br />

to change at all. This switching does not require raising the<br />

potential energy of the particle at any stage,rather the motion<br />

can be powered with energy taken entirely from the thermal<br />

bath by using information about the position of the particle.<br />

This is directly analogous to the mechanism required of<br />

Maxwell s pressure demon (Figure 2b,Section 1.2.1.1),but<br />

does not break the second law of thermodynamics as the<br />

required information transfer (actually,information erasure<br />

[48] ) has an intrinsic energy cost that has to be met<br />

externally.<br />

Figure 10. A type of information ratchet mechanism for transport of a<br />

Brownian particle along a potential-energy surface. [39h, 42b, 45] Dotted<br />

arrows indicate the transfer of information that signals the position of<br />

the particle. If the signal is distance-dependent—say, energy transfer<br />

from an excited state which causes lowering of an energy barrier—<br />

then the asymmetry in the particle’s position between two barriers<br />

provides the “information” which transports the particle directionally<br />

along the potential energy surface.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

It appears to us that information-ratchet mechanisms of<br />

relevance to chemical systems can arise in at least three ways:<br />

1) a localized change to the intrinsic potential-energy surface<br />

depending on the position of the particle (Figure 10); 2) a<br />

position-dependent change in the state of the particle which<br />

alters its interaction with the potential-energy surface at that<br />

point; or 3) switching between two different intrinsic periodic<br />

potentials according to the position of the particle. [39h,49] An<br />

example of the first of these types,in which the system<br />

responds to the “information” from the particle by lowering<br />

the energy barrier to the right-h<strong>and</strong> side (<strong>and</strong> only to the<br />

right-h<strong>and</strong> side) of the particle,is shown in Figure 10.<br />

The particle starts in one of the identical-minima energy<br />

wells (Figure 10a). The position of the particle lowers the<br />

kinetic barrier for passage to the adjacent right-h<strong>and</strong> well <strong>and</strong><br />

it moves there by Brownian motion (10 b!10 c). At this point<br />

it can sample two energy wells by Brownian motion,<strong>and</strong> a<br />

r<strong>and</strong>om reinstatement of the barrier has a 50% chance of<br />

returning the particle to its starting position <strong>and</strong> a 50%<br />

chance of trapping it in the newly accessed well to the right<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

83


Reviews<br />

(Figure 10 d,of course a more efficient mechanism could be<br />

envisaged in which a second information transfer signals the<br />

occupation of the right-h<strong>and</strong> well <strong>and</strong> raises the barrier at this<br />

point only). The particle can no longer go back to the starting<br />

well but is now allowed access to the next well further to the<br />

right. Even though no enthalpic driving force ever exists for<br />

the particle to move from left to right,it is inevitably<br />

transported in that direction through such a mechanism. The<br />

potential-energy surface in Figure 10 is asymmetric,but this is<br />

not necessary if some other means can be devised of<br />

communicating the position of the particle (<strong>and</strong> therefore<br />

breaking spatial inversion symmetry).<br />

Both energy ratchets <strong>and</strong> information ratchets could<br />

provide valid operating mechanisms for synthetic molecularlevel<br />

machines. Furthermore,hybrid mechanisms appear to<br />

be possible. For example,by using asymmetry in the<br />

Brownian particle (for example,a cyclodextrin) rather than<br />

in the potential-energy surface to induce directionality of<br />

motion. [50] Indeed,it seems that the realization of functioning<br />

synthetic chemical machine systems could have as much to<br />

offer theoretical non-equilibrium statistical physics as the<br />

other way around.<br />

1.4.3. From <strong>Motors</strong> to Driving Non-equilibrium Changes in<br />

Chemical Systems<br />

Theoretical fluctuation-driven transport mechanisms have<br />

generally been considered in terms of either two-minima<br />

potential-energy surfaces or an infinitely repeating potentialenergy<br />

surface. The former can be the basis for molecular<br />

switches,while the latter can be employed directly as designs<br />

for molecular-level motors. [51] What other general ideas can<br />

we glean from how these systems work from the viewpoint of<br />

designing other types of synthetic molecular-level machines<br />

that exploit Brownian motion?<br />

1) Asymmetry in some part of the fluctuation-driven transport<br />

cycle is necessary to ensure that time-reversal<br />

symmetry is broken,which means that the particles<br />

undergo nonreciprocal motion during the operating<br />

cycle,thus generating a directional flux. This means that<br />

the particles are not subjected to the opposite of the initial<br />

transport process as the potential is being reset to begin<br />

another cycle.<br />

2) In all these mechanisms (for example,Figures 6–10) we<br />

can see that the particles are always under the influence of<br />

the potential-energy surfaces responsible for transporting<br />

them. Even in the on–off ratchet (Figure 6),the sawtooth<br />

potential can only be switched off for a short time. If it<br />

were switched off for long enough for the particles to<br />

reach an equilibrium distribution over the surface,then<br />

the mechanism would not be able to directionally transport<br />

the particles progressively,it could only switch<br />

between one (average) equilibrium distribution <strong>and</strong><br />

another. In other words,for a machine to operate by<br />

these types of mechanisms which manipulate non-equilibrium<br />

states,Brownian substrates must remain kinetically<br />

associated with the machine throughout its operation—not<br />

necessarily in physical contact but rather not<br />

able to exchange with substrates in the bulk,which are not<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

under the influence of the machine. (This is why we rule<br />

out host–guest complexes that are under thermodynamic<br />

control acting as such machines (Section 1.1).) Exchanging<br />

with the bulk while entering the machine at one end<br />

<strong>and</strong> when leaving at the other is allowed when the two<br />

bulk regions are themselves otherwise separated (for<br />

example,a transmembrane pump).<br />

3) The particles in all fluctuation-driven transport mechanisms<br />

can be considered to be “compartmentalized”,in<br />

that at any moment the particles are free to move only<br />

over a localized area (which continually changes during<br />

the operating cycle of the machine). The compartment<br />

boundaries are not necessarily physical ones; when the<br />

sawtooth potential in the on–off ratchet in Figure 6 is<br />

switched off,it is the short time before the potential is reapplied<br />

that prevents the particles coming to global<br />

equilibrium over the surface.<br />

4) It is the manipulation of the Brownian particles in terms of<br />

which <strong>and</strong> when compartments are “linked” (able to<br />

exchange particles) <strong>and</strong> the statistical balance of the<br />

populations of linked compartments with respect to the<br />

potential-energy surface (in the case of energy ratchets) or<br />

information (in the case of information ratchets) that<br />

enables the transport mechanism to function.<br />

5) Energy ratchets (in which the Brownian substrate is<br />

passive) have been the most widely considered fluctuation-driven<br />

transport mechanisms in the physics literature.<br />

However,the components of chemical systems are not<br />

always passive. For molecular structures it seems likely<br />

that information ratchets (where the position of the<br />

Brownian particle or fragment affects the potentialenergy<br />

surface) will also prove of great utility <strong>and</strong><br />

importance.<br />

6) Fluctuation-driven transport mechanisms are often considered<br />

in terms of charged particles,with electric fields<br />

being applied,typically to generate an infinitely repeating<br />

potential-energy surface. In chemical systems,noncovalent<br />

interactions <strong>and</strong> binding interactions can provide the<br />

necessary changing potential-energy surface,which can be<br />

of limited length <strong>and</strong> not uniform,with the Brownian<br />

substrate being either another ion or molecule or it could<br />

be another fragment of the same molecule. [52]<br />

7) Although the mechanisms for fluctuation-driven transport<br />

have been developed to direct the flow of particles in a given<br />

direction, we can consider applying the same sorts of ideas<br />

to generate other sorts of non-equilibriumdistributions<br />

within molecular-level structures. It should be possible to<br />

design molecular-level machine systems that are neither<br />

switches nor motors by utilizing small numbers of real or<br />

virtual “compartments” within a molecular (or supramolecular)<br />

structure that can be “linked” (<strong>and</strong><br />

“unlinked”) <strong>and</strong> “balanced” (<strong>and</strong> “unbalanced”) in particular<br />

ways,perhaps by using Boolean logic operations,<br />

<strong>and</strong> addressed through orthogonal inputs to perform some<br />

function (see,for example,the “sorting <strong>and</strong> separating”<br />

machine suggested in Section 4.4). Fluctuation-driven<br />

transport mechanisms are illustrative of the general<br />

principles through which non-equilibrium distributions<br />

of supramolecular,molecular,<strong>and</strong> submolecular struc-<br />

84 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

tures can be created,controlled,ordered,<strong>and</strong> manipulated<br />

by inputs of energy.<br />

These are the principles used to develop the concepts for<br />

the compartmentalized molecular machines described in<br />

Section 4.4. However,it is interesting to consider that they<br />

are equally relevant for the design of interacting chemical<br />

reaction cascades <strong>and</strong> catalytic cycles that operate far from<br />

equilibrium.<br />

If biology,mathematics,<strong>and</strong> physics provide the inspiration<br />

<strong>and</strong> strategies for controlling molecular-level motion,it is<br />

through chemistry that artificial molecular-level machine<br />

mechanisms must be designed,constructed,<strong>and</strong> made to<br />

work. The minimum requirements for such systems must be<br />

the restriction of the 3D motion of the machine components<br />

<strong>and</strong>/or the substrate as well as a change in their relative<br />

positions induced by an input of energy. Ways of achieving<br />

control over conformational dynamics involving single bonds<br />

<strong>and</strong> double bonds,as well as over co-conformational dynamics<br />

in kinetically stable supramolecular <strong>and</strong> mechanically<br />

bonded systems,are discussed in Sections 2–7.<br />

2. Controlling Motion in Covalently Bonded<br />

<strong>Molecular</strong> Systems<br />

2.1. Controlling Conformational Changes<br />

2.1.1. Correlated Motion through Nonbonded Interactions<br />

In compounds where two or more bulky planar substituents<br />

are geminally or vicinally connected,the congested steric<br />

environment often dem<strong>and</strong>s a helical ground state where all<br />

the rings are tilted in the same direction,thus generating a<br />

structure reminiscent of the macroscopic screw propellers<br />

found on aeroplanes <strong>and</strong> boats. The lowest energy rotational<br />

processes in such systems generally involve correlated motion<br />

of the planar “blades” (known as “cogwheeling”),<strong>and</strong> their<br />

investigation played an important early role in illustrating<br />

how motion can be controlled by molecular structure. [53]<br />

Inspired by the work of Ōki on the high threefold torsional<br />

barrier in bridgehead-substituted triptycenes, [54] the research<br />

groups of Mislow <strong>and</strong> Iwamura independently replaced the<br />

twofold aryl “rotators” [1k] of molecular propellers with 9triptycyl<br />

units,thereby creating so-called “molecular<br />

gears”. [53c,55–57] In these systems,the blades of each triptycyl<br />

group are tightly intermeshed,so that correlated disrotatory<br />

motion is strongly preferred. Such “dynamic gearing” mirrors<br />

the operation of a macroscopic bevel gear, [58] with correlated<br />

conrotation or uncorrelated rotation amounting to gear<br />

slippage. This threshold mechanism maintains the relationship<br />

between torsional angles of the two rotators. Changing<br />

this relationship can only occur by one of the higher energy<br />

rotational processes,so that stereoisomerism arises when at<br />

least one blade of each triptycyl is differentiable (illustrated<br />

for 1,Scheme 1a). [59] As the isomers differ in the phase<br />

relationship between the substituted rings,the term “phase<br />

isomers” was coined for this subset of residual isomerism. [56c]<br />

Experimental <strong>and</strong> theoretical studies confirm that the correlated<br />

disrotation is generally favored over other torsional<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Scheme 1. a) The three residual diastereoisomers of molecular gear<br />

1. [57b] Correlated disrotation maintains the phase relationship between<br />

the labeled blades (shown here in red) for each isomer. Interconversion<br />

between isomers (“gear slippage”) requires correlated conrotation<br />

or uncorrelated rotation. b) <strong>Molecular</strong> gear train 2. [60] c) Vinyl molecular<br />

gear 3 (shown as the racemic residual diastereomer); [61] a<br />

molecular gear 4 with a 2:3 gearing ratio; [62j] <strong>and</strong> a molecular gear 5<br />

based on revolution around a metallocene <strong>and</strong> with a 3:4 gearing<br />

ratio. [63] The arrows show correlated motions but are not meant to<br />

imply intrinsic directionality.<br />

processes by 30–40 kcal mol 1<br />

in these systems,largely<br />

because of the remarkable ease of the disrotary motion<br />

(DG ° = ca. 1–2 kcalmol 1 ). This result actually represents<br />

stricter selectivity than occurs for the selection rules that<br />

govern correlated torsions under orbital symmetry control<br />

(see Section 2.1.5).<br />

Iwamura <strong>and</strong> co-workers successfully extended the<br />

dynamic gearing concept to multiple gear systems in which<br />

each adjacent pair must disrotate,so that the behavior of the<br />

outermost rotators is dependent of the number of linkages in<br />

the chain. In 2 therefore (Scheme 1b),despite the large<br />

number of possible conformations <strong>and</strong> increased flexibility of<br />

the chain,the two outer triptycyl units conrotate <strong>and</strong> the<br />

phase relationship of the two chlorine labels is strictly<br />

maintained in both phase isomers. [60]<br />

Dynamic gearing has also been realized in vicinally linked<br />

triptycyl rotors,such as 3 (Scheme 1c), [61] <strong>and</strong> in systems with<br />

gearing ratios other than 3:3,for example,2:3 (4), [62] <strong>and</strong> 3:4<br />

(5). [63] Although the low-energy dynamic processes observed<br />

in 5 could not be unequivocally shown to be correlated,this<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

85


Reviews<br />

system is one of the first<br />

molecular structures in<br />

which the facile rotation<br />

of p fragments in metallocenes<br />

was utilized,thus<br />

prompting analogy with a<br />

ball bearing. [64]<br />

The same comparison<br />

can be applied to a more<br />

recent series of trinuclear<br />

s<strong>and</strong>wich complexes in<br />

which multiple correlated<br />

motions have been demonstrated.<br />

[65] Each silver<br />

cation in [Ag 3(6) 2] has a<br />

linear coordination geometry,bound<br />

to one nitrogen<br />

donor from each trismonodentate<br />

disk-shaped<br />

lig<strong>and</strong> (red)—the resulting<br />

complex is helical,<br />

with all the thiazolyl<br />

rings tilted in the same<br />

direction (Figure 11b).<br />

R<strong>and</strong>om helix inversion<br />

is a low-energy process,<br />

which occurs through a<br />

nondissociative “flip”<br />

motion whereby rotation<br />

of the heteroaromatic<br />

rings so that they point<br />

in the opposite direction<br />

is accompanied by a 1208<br />

relative rotation of the two large disks (illustrated for<br />

[Ag 3(6)(7)] in Figure 11 c). [65a] In the heteroleptic analogue<br />

[Ag 3(6)(7)],a similar coordination geometry is adopted so<br />

that only every alternate thiazolyl ring in the hexakismonodentate<br />

disk lig<strong>and</strong> (blue) is bound at any one time.<br />

The “flip” helix reversal,during which the lig<strong>and</strong> <strong>and</strong> metal<br />

partners do not change,can clearly still occur. Lig<strong>and</strong><br />

exchange is also rapid,however,<strong>and</strong> most likely occurs via<br />

the trigonal transition state illustrated in Figure 11 c. The<br />

overall result is a correlated rotation of the heteroaromatic<br />

rings together with a 608 relative rotation of the two disks. If a<br />

lig<strong>and</strong> exchange <strong>and</strong> a flip step occur concurrently,the sense<br />

of rotation in each step is opposite,so that overall a 608<br />

relative rotation of the disk-shaped lig<strong>and</strong>s occurs. [65b,c]<br />

While all these <strong>and</strong> other [66] related studies clearly<br />

demonstrate the role steric interactions can play,at equilibrium<br />

the submolecular motions are nondirectional even<br />

within a partial rotational event. Simply restricting the<br />

thermal rotary motion of one unit by a larger blocking<br />

group or by the similarly r<strong>and</strong>om motion of another unit<br />

cannot,in itself,lead to directionality. A molecular machine<br />

requires some form of external modulation over the dynamic<br />

processes to drive the system away from equilibrium <strong>and</strong><br />

break detailed balance.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 11. a) Chemical structure of tris-monodentate disk-shaped lig<strong>and</strong> 6 <strong>and</strong> hexakis-monodentate lig<strong>and</strong><br />

7. [65] b) Schematic representation of complex [Ag 3(6) 2] arbitrarily shown as its M-helical enantiomer.<br />

c) Description of the two correlated rotation processes occurring in [Ag 3(6)(7)]. The direction of rotation for<br />

an M!P transition through lig<strong>and</strong> exchange is opposite to that for the potentially subsequent P!M’<br />

transition by the nondissociative flip mechanism. Schematic representations reprinted with permission from<br />

Refs. [65a,c].<br />

2.1.2. Stimuli-Induced Conformational Control around a Single<br />

Covalent Bond<br />

As a first step towards achieving controlled <strong>and</strong> externally<br />

initiated rotation around C C single bonds,Kelly <strong>and</strong> coworkers<br />

combined triptycene structures with a molecularrecognition<br />

event. [67] In the resulting “molecular brake” 8<br />

Scheme 2. “<strong>Molecular</strong> brakes” induced by a) metal-ion binding [68] <strong>and</strong><br />

b) redox chemistry. [70] EDTA = ethylenediaminetetraacetate, mCPBA =<br />

meta-chloroperbenzoic acid.<br />

86 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

(Scheme 2a), [68] free rotation of a triptycyl group is halted by<br />

the conformational change brought about by complexation of<br />

the appended bipyridyl unit with Hg 2+ ions,thus effectively<br />

putting a “stick” in the “spokes”. [69] A similar “braking effect”<br />

has been demonstrated by using redox chemistry <strong>and</strong> Narylindolinone<br />

9 (Scheme 2b); both oxidized forms of the<br />

sulfur side chain exhibit markedly increased barriers to<br />

rotation around the N aryl bond (see<br />

also 45,Scheme 21). [70,71]<br />

Kelly et al. next extended their<br />

investigation of restricted Brownian<br />

rotary motion to a molecular realization<br />

(10) of the Feynman adiabatic<br />

ratchet <strong>and</strong> pawl, [72] in which a helicene<br />

plays the role of the pawl in attempting<br />

to direct the rotation of the attached<br />

triptycene “cog wheel” in one direction<br />

as a result of the chiral helical structure.<br />

Although the calculated energetics for<br />

rotation showed an asymmetric potential<br />

energy profile (Figure 12 b),<br />

NMR experiments confirmed that rotation<br />

occurred with equal frequency in<br />

both directions. This result is,of course,<br />

in line with the conclusions of the<br />

Feynman thought experiment (Section<br />

1.2.1.4). [73] The rate of a molecular<br />

transformation (clockwise <strong>and</strong> anticlockwise<br />

rotation in 10 included)<br />

depends on the energy of the transition<br />

state (<strong>and</strong> the temperature),not the<br />

shape of the energy barrier: state<br />

functions such as enthalpy <strong>and</strong> free<br />

energy do not depend on a system s<br />

history. [74] Thus,although rotation in 10<br />

follows an asymmetric potential-energy surface,the principle<br />

of detailed balance at equilibrium requires that transitions in<br />

each direction occur at equal rates. [75]<br />

The essential element missing from 10 needed to turn the<br />

triptycene directionally is some form of energy input to drive<br />

it away from equilibrium <strong>and</strong> break the detailed balance. In<br />

principle,this could be achieved rather simply for 10 by a<br />

periodic fluctuation in temperature (causing directional<br />

Figure 12. a) Kelly’s molecular realization (10) of Feynman’s adiabatic<br />

ratchet <strong>and</strong> pawl which does not rotate directionally at equilibrium. [72]<br />

b) Schematic representation of the calculated enthalpy changes for<br />

rotation around the single degree of internal rotational freedom in 10.<br />

1 H<br />

Angew<strong>and</strong>te<br />

Chemie<br />

rotation in chemical structures does not have to be complicated),thus<br />

constituting a temperature ratchet. [75] However,<br />

proving this experimentally by quantifying the net rotation<br />

appears to be nontrivial. As a different solution,Kelly et al.<br />

proposed a modified version of the ratchet structure (11a,<br />

Scheme 3),in which a chemical reaction is used as the source<br />

of energy. [76] If the amino group is ignored,all three energy<br />

Scheme 3. A chemically powered unidirectional rotor. [76] Priming of the rotor in its initial state with<br />

phosgene (11a!12 a) allows a chemical reaction to take place when the helicene rotates far enough up<br />

its potential well towards the blocking triptycene arm (12b). This gives a tethered state 13 a, for which<br />

rotation over the barrier to 13 b is an exergonic process that occurs under thermal activation. Finally, the<br />

urethane linker can be cleaved to give the original molecule with the components rotated by 1208 with<br />

respect to each other (11b).<br />

minima for the position of the helicene with respect to the<br />

triptycene “teeth” are identical: the energy profile for 3608<br />

rotation would appear as three equal energy minima,<br />

separated by equal barriers. As the helicene oscillates back<br />

<strong>and</strong> forth in a trough,however,sometimes it will come close<br />

enough to the amine for a chemical reaction to occur (as in<br />

12b). Priming the system with a chemical “fuel” (phosgene in<br />

this case to give the isocyanate 12a) results in “ratcheting” of<br />

the motion some way up the energy barrier (13a). Continuation<br />

of the rotation in the same direction,over the energy<br />

barrier,can occur under thermal control <strong>and</strong> is now an<br />

exergonic process (giving 13b) before cleavage of the<br />

urethane gives the system rotated by 1208 (11 b). Although<br />

the current version of the system can only carry out one third<br />

of a full rotation,it demonstrates the principles required for a<br />

fully operating <strong>and</strong> cyclable rotary system under chemical<br />

control <strong>and</strong> represents a major advance in the experimental<br />

realization of molecular-level machines.<br />

Exemplified by the motor proteins from nature,continually<br />

operating (autonomous) chemically powered molecular<br />

motors may be classified as a subset of catalysts—catalyzing<br />

the conversion of high-energy “fuel” molecules into lower<br />

energy products while undergoing a conformational change<br />

but returning to the starting state on completion of each cycle.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

87


Reviews<br />

Mock <strong>and</strong> Ochwat have proposed 14,which undergoes the<br />

catalytic cycle illustrated in Scheme 4,as a minimalist model<br />

of a molecular motor. [77] In the process of catalyzing the<br />

hydration of its ketenimine fuel (red box) to give a<br />

Scheme 4. Proposed system for chemically driven directional 1808<br />

rotary motion fueled by hydration of a ketenimine (red box). [77] For<br />

clarity, the catalytic cycle for a single enantiomer of motor 14 is shown.<br />

This cycle produces the opposite enantiomer, which undergoes an<br />

analogous cycle in which rotation must be biased in the opposite<br />

direction. A number of potential side reactions (not illustrated) which<br />

would divert the system from the catalytic cycle were found to be<br />

kinetically insignificant under the reaction conditions employed.<br />

carboxamide waste product (blue box),epimerization of the<br />

stereogenic center in the motor occurs—formally rotation<br />

around a C O bond (as indicated for (R)-14 in Scheme 4).<br />

The cycle operates (Scheme 4) under the specific kinetic<br />

conditions k 2,k 4 > k 1[AcCH=C=NtBu] > k 3[H 2O]. Kinetic<br />

analysis based on spectrophotometric measurements was<br />

used to identify reaction conditions under which this relation<br />

is satisfied. Over a single catalytic cycle,therefore,the<br />

molecular chirality should ensure some biasing of the rotational<br />

direction for formal 1808 rotation around the C O<br />

bond. However,the subsequent cycle on the opposite<br />

enantiomer will experience the antipodal directing effect so<br />

that the biasing effect is the exact reverse to before. [78]<br />

The restriction of rotational motion arising from steric<br />

interactions between ortho substituents in biphenyl systems<br />

forms the basis for another system designed to exhibit<br />

directional rotation through ring opening/ring closing of a<br />

chiral lactone (Scheme 5). [79] Ring opening of lactone 15 with<br />

a bulky nucleophile should proceed preferentially through<br />

attack at one of the two diastereotopic faces (k 15!16-aS ><br />

k 15!16-aR). Equilibration of the ring-opened diastereomers<br />

should then occur by r<strong>and</strong>om oscillations around the aryl–aryl<br />

bond,with the bulky nucleophile preferentially avoiding<br />

passing over the cyclohexanol ring. Ring closing is then<br />

proposed to be faster from the 16-aR isomer because of the<br />

proximity of the reacting groups,so that a net 3608 rotation<br />

will be accomplished by a number of molecules in the system.<br />

Experimentally,equilibration of the two ring-opened diastereomers<br />

in the current system occurs too fast for determination<br />

of the selectivity of the ring-opening reaction. Similarly,demonstrating<br />

directional selectivity for the ringclosing<br />

step will present a significant challenge. A further<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Scheme 5. Proposed system for unidirectional rotation around an aryl–<br />

aryl bond based on ring opening/ring closing of a chiral lactone. [79]<br />

Although this cannot function at equilibrium, it could if additional<br />

features were added such that k 15!16-aS ¼6 k 16-aS!15 <strong>and</strong> k 15!16-aR ¼6<br />

k 16-aR!15.<br />

modification is also necessary to achieve directional 3608<br />

rotation; since the ring-closing step is the exact reverse of the<br />

ring-opening step,both must proceed by the same transition<br />

state so that the rate in either direction is the same (k 15!16-aS =<br />

k 16-aS!15 <strong>and</strong> k 15!16-aR = k 16-aR!15). No energy would be consumed<br />

by the mechanism proposed in Scheme 5 <strong>and</strong>,of<br />

course,the principle of detailed balance (Section 1.4.1)<br />

dem<strong>and</strong>s that unidirectional rotation cannot occur in a<br />

system at equilibrium.<br />

Stereoselective ring opening of racemic biaryl lactones<br />

using chiral reagents results in a directional rotation of about<br />

908,which can be extended to a full half-turn by chemoselective<br />

ring closing to give the lactone by using a hydroxy<br />

group in the other ortho position. This effect has been<br />

demonstrated by Branchaud <strong>and</strong> co-workers [80] <strong>and</strong>,in an<br />

independent effort,Feringa <strong>and</strong> co-workers have successfully<br />

extended this strategy to obtain a full 3608 rotation around a<br />

C C single bond. [81] The process involves four intermediates<br />

(A–D,Figure 13 a),in each of which rotation around the<br />

biaryl bond is restricted: by covalent attachment in A <strong>and</strong> C,<br />

<strong>and</strong> through nonbonded interactions in B <strong>and</strong> D. Directional<br />

rotation to interchange these intermediates requires a stereoselective<br />

bond-breaking reaction in steps (1) <strong>and</strong> (3) <strong>and</strong> a<br />

regioselective bond-formation reaction in steps (2) <strong>and</strong> (4).<br />

Unlike 15 above,lactones 17 <strong>and</strong> 19 (Figure 13 b) exist as<br />

racemic mixtures as a result of a low barrier for small<br />

amplitude rotations around the aryl–aryl bond. Reductive<br />

ring opening with high enantioselectivity is,however,achievable<br />

for either lactone by using a homochiral borolidine<br />

catalyst,<strong>and</strong> the released phenol can subsequently be<br />

orthogonally protected to give 18a or 20 a. The ring-opened<br />

compounds are produced in near-enantiopure form in a<br />

process which involves directional rotation of 908 around the<br />

biaryl bond,governed by the chirality of the catalyst,<strong>and</strong><br />

powered by consumption of borane. The ortho substitution of<br />

these species results in a high barrier to axial rotation.<br />

Oxidation of the benzylic alcohol (18 a!18 b or 20 a!20b)<br />

primes the motor for the next rotational step. Selective<br />

88 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 13. a) Schematic representation of unidirectional rotation around a single bond, through four states. [81] In states A <strong>and</strong> C rotation is<br />

restricted by a covalent linkage, but the allowed motion results in helix inversion. In states B <strong>and</strong> D rotation is restricted by nonbonded<br />

interactions between the two halves of the system (red <strong>and</strong> green/blue). These forms are configurationally stable. The rotation relies on<br />

stereospecific cleavage of the covalent linkages in steps (1) <strong>and</strong> (3), then regiospecific formation of covalent linkages in steps (2) <strong>and</strong> (4).<br />

b) Structure <strong>and</strong> chemical transformations of a unidirectional rotor. Reactions: 1) Stereoselective reduction with (S)-CBS then allyl protection.<br />

2) Chemoselective removal of the PMB group resulting in spontaneous lactonization. 3) Stereoselective reduction with (S)-CBS then protection<br />

with a PMB group. 4) Chemoselective removal of the allyl group resulting in spontaneous lactonization. 5)Oxidation to carboxylic acid.<br />

removal of one of the protecting groups on the enantiotopic<br />

phenols results in spontaneous lactonization when thermally<br />

driven axial rotation brings the two reactive groups<br />

together—again probably a net directional process because<br />

of the steric hindrance of the ortho substituents (although this<br />

is not demonstrated because the chirality is destroyed in this<br />

step). Figure 13b illustrates the unidirectional process achieved<br />

using the (S)-Corey–Bakshi–Shibata ((S)-CBS) catalyst;<br />

rotation in the opposite sense can be achieved by<br />

employing the opposite borolidine enantiomer <strong>and</strong> swapping<br />

the order of phenol protection <strong>and</strong> deprotection steps.<br />

2.1.3. Stimuli-Induced Conformational Control in<br />

Organometallic Systems<br />

Controlling the facile rotary motion of<br />

lig<strong>and</strong>s in metal s<strong>and</strong>wich or double-decker<br />

complexes (introduced in Section 2.1.1) is conceptually<br />

similar to controlling rotation around<br />

covalent single bonds,<strong>and</strong> stimuli-induced<br />

control in such metal complexes has also been<br />

demonstrated. In metal bisporphyrinate complexes<br />

such as [Ce(21)] (Scheme 6) rotary<br />

motion corresponds to enantiomerization<br />

when the lig<strong>and</strong>s possess D 2h symmetry. This<br />

situation provides a convenient h<strong>and</strong>le for<br />

monitoring the kinetics of rotation. [82] In complex<br />

[Ce(21)],rotation around the metal center<br />

is slow enough to permit isolation of the two<br />

enantiomers by chiral HPLC. However,reduc-<br />

Angew<strong>and</strong>te<br />

Chemie<br />

tion of the metal center ([Ce IV (21)]![Ce III (21)] ) increases<br />

the rate of racemization by over 300-fold. [82c] This effect is<br />

thought to derive from a reduced p–p interaction between the<br />

two lig<strong>and</strong>s as a consequence of the larger ionic radius of the<br />

metal center in the lower oxidation state. Similarly,complexes<br />

formed around the smaller Zr IV ion show very slow rotational<br />

dynamics at pH 7, [83] yet protonation results in facile racemization.<br />

[82a] The effect is retarded by oxidation of the<br />

porphyrin lig<strong>and</strong>s,probably as a result of electrostatic<br />

repulsion of incoming protons,but also because of the fact<br />

that the highest occupied molecular orbital (HOMO) of the<br />

complex is antibonding,so removal of electrons should<br />

Scheme 6. Controlling the rate of relative rotations of porphyrin lig<strong>and</strong>s in cerium bis[tetrakisarylporphyrinate]<br />

double-decker complex [Ce(21)]. [82c]<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

89


Reviews<br />

increase the p–p bonding interaction between the two<br />

porphyrin rings. [82c]<br />

Rotary motion in s<strong>and</strong>wich complexes can be interrupted<br />

by the binding of a ditopic guest between recognition sites on<br />

each of the two rings. If more than one such pair of binding<br />

sites exists,a positive,homotropic allosteric effect is<br />

observed. [84] This principle has been exploited to create<br />

allosteric receptors (for example, 22) for g-dicarboxylic<br />

acids (Scheme 7), [85] b-dicarboxylate anions, [86] potassium<br />

Scheme 7. Positive homotropic allosteric binding in a double-decker complex 22. [85]<br />

A result of the cooperative binding is pronounced guest selectivity; for example, ddicarboxylic<br />

acids are not bound. The chirality of the guest is communicated to the host,<br />

thus determining the sense of the axial chirality induced about the metal–porphyrin axis.<br />

ions, [87] silver ions, [88] as well as mono- <strong>and</strong> oligosaccharides<br />

(even in aqueous solution). [89] It was originally proposed that<br />

such effects were primarily a result of the restriction of the<br />

motion of the porphyrin ring when the first guest binds,thus<br />

reducing the entropic cost of subsequent identical binding<br />

events. [85–87,89] In the case of cooperative binding of silver ions,<br />

however,it is observed that binding of the guest species<br />

progressively <strong>and</strong> nonlinearly increases the rate of rotation.<br />

[88b] Here,cation binding induces a deformation of the<br />

porphyrin rings away from planarity to generate a more<br />

domed structure,simultaneously weakening the p–p interactions<br />

between the porphyrin rings <strong>and</strong> preorganizing the<br />

remaining silver binding sites. Recent results suggest that<br />

torsional strain induced by binding may play a significant role<br />

in the cooperative binding in all these systems. [90,91]<br />

Ferrocene has long been known to exhibit a low barrier to<br />

cyclopentadienyl (Cp) ring rotation. [92] Recently,a gas-phase<br />

experimental <strong>and</strong> theoretical study of dianionic 23 2<br />

<strong>and</strong><br />

anionic [23·H] indicated that a two-state rotary switch could<br />

be operated by monoprotonation <strong>and</strong> deprotonation<br />

(Scheme 8). [93] The conformation of the dianion is governed<br />

by coulombic repulsions,while an intramolecular hydrogen<br />

bond stabilizes the observed conformation in the monoanion.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Scheme 8. Proton-switched intramolecular rotation in a ferrocene<br />

derivative. [93] The amplitude of relative rotary motion of the two<br />

cyclopentadienyl lig<strong>and</strong>s is about 1128.<br />

In contrast to these simpler metallocene or bis-<br />

(porphyrinate) examples,metallacarboranes tend to<br />

have rather high barriers to rotation around the metal–<br />

lig<strong>and</strong> axis. Dicarbollide lig<strong>and</strong> 24 2<br />

features two<br />

adjacent carbon atoms on its bonding face which<br />

imparts a dipole moment perpendicular to the metal–<br />

lig<strong>and</strong> axis (Figure 14a). Transoid complexes are<br />

normally preferred to minimize electrostatic repulsion.<br />

[94] Two exceptions,however,are the Ni IV or Pd IV<br />

complexes which are cisoid. [95] Complex [Ni(24) 2] can<br />

therefore operate as a reversible rotary switch on<br />

manipulation of the oxidation state or by photoexcitation<br />

(Figure 14b). [96]<br />

2.1.4. Stimuli-Induced Conformational Control around<br />

Several Covalent Bonds<br />

With an external stimulus,it is possible to control<br />

the conformation around not just one bond,but<br />

several all at once. Conformational control in biopolymers<br />

is of great importance in structural biology,<strong>and</strong><br />

a vast array of methods for artificially triggering,<br />

altering,<strong>and</strong> reversing folding in polypeptides <strong>and</strong><br />

polysaccharides now exists. [97] In nature,as in artificial<br />

systems,host–guest binding is often used to trigger<br />

complex conformational changes in one or both of the<br />

interacting species. In some cases this involves no more than<br />

minor bond deformations or the restriction of a few rotational<br />

degrees of freedom. In others,molecular recognition results<br />

in significant changes to a single degree of freedom (see<br />

Section 2.1.3),yet binding of one chemical species to another<br />

can also often result in significant changes to the conforma-<br />

Figure 14. a) Dicarbollide lig<strong>and</strong> 24 2 . b) Metallacarborane [Ni(24) 2]as<br />

an electrochemically controlled rotary switch. [96] The cisoid–transoid<br />

exchange involves a rotation of 1448. Both formal reduction <strong>and</strong><br />

photochemical excitation of [Ni IV (24) 2] populate the lowest unoccupied<br />

molecular orbital (LUMO) <strong>and</strong> therefore elicit similar dynamic<br />

responses. Black spheres represent boron atoms, white spheres<br />

carbon atoms.<br />

90 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

tion <strong>and</strong> internal motion around several bonds. Such<br />

“induced fit” mechanisms are the basis of many<br />

allosteric systems,whereby recognition of one species<br />

affects the binding properties or enzymatic activity at a<br />

remote site in the same molecule. Allosterism,cooperativity,<strong>and</strong><br />

feedback are central to many functional<br />

biological systems [98] <strong>and</strong> synthetic approaches towards<br />

achieving similar effects have been extensively reviewed<br />

elsewhere. [84,99] Here we are interested primarily in the<br />

dynamic processes themselves <strong>and</strong>,therefore,in many<br />

cases where the conformational changes are rather small<br />

or poorly defined,such systems do not fall under our<br />

definition of molecular machines. However,certain<br />

examples do exhibit impressive control over the molecular<br />

shape or conformation <strong>and</strong> merit discussion here in<br />

their own right,together with other examples of stimuliinduced<br />

conformational changes around several bonds.<br />

Some of the first <strong>and</strong> most elegant examples of<br />

synthetic allosteric receptors were introduced by Rebek<br />

et al. (Scheme 9). [99a,100] In 25,the binding of a metal ion<br />

such as tungsten to the bipyridyl lig<strong>and</strong> forces coplanarity<br />

of this unit,thus distorting the crown ether<br />

conformation <strong>and</strong> diminishing its ability to bind potas-<br />

Scheme 9. a) Negative heterotropic allosteric receptor 25 binds alkali metal ions, with selectivity<br />

for K + . [100] Chelation of tungsten to the bipyridyl moiety forces this unit to adopt a rigid<br />

conformation in which the 3 <strong>and</strong> 3’ substituents are brought close together. The resulting<br />

conformation of the crown ether does not favor binding through all the oxygen atoms <strong>and</strong> so<br />

affinity for K + ions is reduced. In fact [W(CO) 3(25)] shows a preference for binding the smaller<br />

Na + ion. b) Positive homotropic allosteric receptor 26. [101]<br />

sium ions—a negative heterotropic allosteric effect. The same<br />

research group later developed this concept to create positive,<br />

homotropic allosteric receptor 26,in which binding one<br />

molecule of Hg(CN) 2 preorganizes the remaining binding site<br />

for a second binding event. [101] These strategies have since<br />

either directly spawned or indirectly inspired an extraordinarily<br />

wide range of increasingly sophisticated synthetic<br />

allosteric receptors. [84,99]<br />

Binding to molecular tweezer <strong>and</strong> clip receptors<br />

(Figure 15) can result in significant conformational changes<br />

in the host,as favorable interactions with the guest are<br />

maximized. In 27,for example,the distance between the<br />

sidewalls decreases from 14.5 to 6.5 Š on binding to aromatic<br />

guests, [102] while diphenylglycouril-based clips such as 28 do<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Figure 15. Binding-induced conformational changes in molecular clips. a) Dimethylenebridged<br />

aromatic clip 27 in which binding to 1,2,4,5-tetracyanobenzene results in a<br />

large-amplitude induced-fit conformational change (motion indicated by arrows). [102]<br />

b) The three accessible conformations of the diphenylglycouril-based clips illustrated<br />

for simple derivative 28 (methoxy substituents are not shown on the space-filling<br />

models). [103] The syn,anti (sa) form dominates in solution, but only the anti,anti form is<br />

able to bind aromatic guests—in an induced-fit mechanism. c) Glycouril-based clip 29<br />

in which an additional binding site (a crown ether) can control the conformation of the<br />

clip molecule, preorganizing the p-electron-rich binding site <strong>and</strong> resulting in a positive,<br />

heterotropic allosteric effect. [104] The space-filling representations are reprinted with<br />

permission from Ref. [103a].<br />

not principally adopt the anti,anti (aa)<br />

binding conformation,only doing so in<br />

the presence of a suitable guest. [103]<br />

Similar structures can incorporate two<br />

conformationally coupled binding sites<br />

<strong>and</strong> exhibit allosteric effects. Glycourilbased<br />

clip 29 bears two crown ether<br />

substituents <strong>and</strong> binds potassium ions<br />

to form a complex with 1:2 stoichiometry.<br />

The binding event stabilizes the<br />

aa conformation,preorganizing the<br />

electron-rich naphthalene units in a<br />

pseudoparallel arrangement,thus<br />

increasing the binding affinity for<br />

simple electron-poor aromatic compounds<br />

such as 1,3-dinitrobenzene. [104]<br />

In atropoisomeric host molecules,an<br />

induced-fit binding process at elevated<br />

temperatures can be used to dynamically<br />

select the single optimal binding<br />

conformation for a particular guest,<br />

which can then be “saved” by cooling the system to ambient<br />

temperature <strong>and</strong> removal of the templating guest. [105]<br />

The addition of an external lig<strong>and</strong> or guest can also be<br />

used to disrupt a preexisting intramolecular interaction <strong>and</strong><br />

cause a change in conformation. This is the case for 4,4’bipyridyl-capped<br />

zinc porphyrin 30 in which the pyridyl <strong>and</strong><br />

porphyrin planes are switched from perpendicular to mutually<br />

parallel on addition of pyridine (Scheme10). [106]<br />

It is well known that the relative stabilities of the two<br />

different chair forms of six-membered alicyclic rings can be<br />

manipulated by covalent substitutions. [107,108] However,in<br />

trisaccharide 31 this conformational change can be achieved<br />

by adding <strong>and</strong> removing metal ions (Scheme 11). [109] In the<br />

absence of metal ions,the 4 C 1 conformation is preferred for<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

91


Reviews<br />

Scheme 10. Conformational switching in capped porphyrin 30. [106]<br />

Scheme 11. Chelation-induced conformational control in a<br />

trisaccharide. [109b]<br />

the central “hinge” pyranose unit,in which all the substituents<br />

are equatorial <strong>and</strong> both ends of the molecule are far apart. On<br />

addition of Pt II ions,however,a ring flip to the 1 C 4 form is<br />

observed so that the amine units can chelate to the metal ion<br />

<strong>and</strong> results in a major change in the shape of the oligosaccharide.<br />

Removal of the metal ions with NaCN returns the<br />

trisaccharide to its original extended conformation. Such<br />

stimuli-induced conformational switching has been achieved<br />

in a variety of cyclic systems through binding of metal ions [110]<br />

<strong>and</strong> variation of the pH value. [111] In addition to having<br />

significance for the control <strong>and</strong> modeling of biologically<br />

relevant oligosaccharides, [112] such systems have served as<br />

inspiration for the development of synthetic conformational<br />

switches (see Section 8.2.1).<br />

Well-defined large amplitude (sometimes very large<br />

amplitude; see Section 8.2.1) conformational changes can be<br />

induced in cavit<strong>and</strong>s derived from resorcin[4]arenes (for<br />

example, 32,Scheme 12). [113] These molecules can exist in<br />

either an open “kite” conformation or a narrower “vase”<br />

form. As a result of solvation effects,the kite form is generally<br />

favored at low temperatures <strong>and</strong> the vase form at high<br />

temperatures. [114] Recently,it has been demonstrated that<br />

reversible switching between the two can also be achieved at<br />

room temperature through protonation of the quinoxaline<br />

nitrogen atoms, [115] or coordination of metal ions to these<br />

same atoms, [116] while the switching can also be highly<br />

dependent on the solvent. [115b,c] In a related series of cavit<strong>and</strong>s<br />

bearing amide substituents around the rim,the kite form is<br />

stabilized by dimerization (so-called “velcr<strong>and</strong>s” forming<br />

“velcraplexes”),which is driven by intermolecular hydrogen-<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Scheme 12. Conformational bistability in resorcin[4]arene cavit<strong>and</strong>s 32.<br />

Reversible switching between the two forms can be achieved with a<br />

number of stimuli: a) by changing the temperature; for example, for<br />

R = Me in CDCl 3/CS 2:1)T < 211 K; 2) T > 318 K; [114] b) by changing the<br />

pH value; for example, for R = 3,5-di(tert-butyl)phenylmethyl in CDCl 3<br />

at T = 298 K: 1) CF 3COOH; 2) K 2CO 3; [115] c) by adding metal ions; for<br />

example, for R = n-Hex in CDCl 3 at 295 K: 1) ZnI 2 in [D 4]methanol;<br />

2) excess [D 4]methanol. [116]<br />

bonding,van der Waals,<strong>and</strong> solvophobic forces. Increasing<br />

the temperature or solvent polarity,as well as the introduction<br />

of suitable guest species,can induce switching to the monomeric<br />

vase forms. [117] Conversely,the use of amide substituents<br />

that are able to stabilize the vase form through intramolecular<br />

hydrogen bonding results in kinetically stable<br />

inclusion complexes in the presence of a suitable guest. [118]<br />

Incorporation of metal-ligating carboxymethylphosphonate<br />

groups into these structures allows switching of the preferred<br />

conformation from vase to kite,along with concomitant<br />

release of any encapsulated guest,on addition of lanthanum<br />

ions. [119] The reverse process occurs on addition of a competing<br />

lig<strong>and</strong> for the metal ion. The calixarene family,in general,<br />

clearly share this potential for interesting conformational<br />

dynamics, [120] <strong>and</strong> in certain cases significant degrees of<br />

control over these can be achieved,thus resulting in a<br />

number of induced-fit <strong>and</strong> allosteric receptor systems. [121]<br />

Rather than adding an external species to induce a<br />

conformational change on binding,intramolecular interactions<br />

between two parts of the same molecule can be<br />

controlled remotely to bring about changes in the geometry.<br />

Macrocyclic compound 33 4+ preferentially adopts the selfcomplexed<br />

conformation shown in Scheme 13,which is<br />

stabilized both by p–p charge-transfer interactions <strong>and</strong> by<br />

hydrogen bonds. [122] Electrochemical reduction of the cyclophane,however,“switches<br />

off” these interactions <strong>and</strong> the<br />

molecule adopts a (poorly defined) conformation in which the<br />

dioxynaphthalene unit is no longer encapsulated by the<br />

macrocycle. [123]<br />

The formation of excimers or exciplexes between two<br />

chromophores in the same molecule is another way by which<br />

intramolecular interactions may be remotely controlled so as<br />

to effect a conformational change. Most examples,however,<br />

involve flexible,linear ground states with poorly defined<br />

geometries. [124] In certain donor–bridge–acceptor systems,a<br />

conformational change does not occur in the initial photochemically<br />

excited state,rather,electron transfer occurs to<br />

give an “extended charge transfer” species. Depending on the<br />

92 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 13. Electrochemically induced conformational motion based on remote control of intramolecular<br />

interactions. [122]<br />

rigidity of the structure,electrostatic attraction<br />

can then drive a conformational change to a<br />

“compact charge transfer” state. Such systems<br />

are often referred to as “molecular harpoons”.<br />

[125]<br />

In recent years,a number of oligomeric<br />

synthetic systems have been investigated in<br />

solution for their ability to adopt well-defined<br />

conformations with interesting <strong>and</strong> tunable<br />

secondary <strong>and</strong> tertiary structures,particularly<br />

those of a helical nature. [126] Oligomeric phenylacetylenes,for<br />

example,form helical structures<br />

in polar solvents as a result of solvophobic<br />

interactions <strong>and</strong> thus exhibit solvent- <strong>and</strong><br />

temperature-sensitive secondary structures.<br />

[127,128] Aromatic oligoamide foldamers,<br />

on the other h<strong>and</strong>,adopt helical conformations<br />

in nonpolar solutions,<strong>and</strong> in the solid state,<br />

given suitably arranged hydrogen-bonding<br />

substituents. [126h-j] For heptamer 34,helix formation<br />

is driven by intramolecular hydrogen<br />

bonding between the amide groups <strong>and</strong> pyridine<br />

units (Figure 16 a). [129] Under certain conditions,double<br />

helices (in which two str<strong>and</strong>s of<br />

the same oligomeric monomer are intertwined)<br />

are formed. This structure is mainly<br />

stabilized through intermolecular p–p stacking<br />

interactions,with hydrogen bonding determining<br />

the helical conformation of each str<strong>and</strong>.<br />

The double-helix structure is asymmetric,with<br />

each end of a str<strong>and</strong> in different environments. Interconversion<br />

of two equivalent forms occurs by a rotational corkscrew<br />

motion of one str<strong>and</strong> relative to the other. Significant heating<br />

is required to cause the much larger conformational changes<br />

required to bring the system into the single-helix regime<br />

(Figure 16 b). [129a,c]<br />

In closely related heptamer, 35,the single-helix secondary<br />

structure is disrupted upon protonation of the four diaminopyridine<br />

units,thus giving an extended linear conformation<br />

(35·(H + ) 4,Scheme 14). [130] Further protonation of the three<br />

remaining dicarboxypyridine moieties yields a second helical<br />

arrangement (35·(H + ) 7),which on deprotonation can be<br />

returned through the linear tetraprotonated form back to<br />

the neutral helix. A pentadecamer analogue has recently been<br />

shown to undergo a similar reversible uncoiling/coiling transformation,which<br />

results in a remarkable change in the overall<br />

Angew<strong>and</strong>te<br />

Chemie<br />

length from 12.5 to 57 Š. [131] Using<br />

the same principles of altering<br />

hydrogen-bonding arrangements,<br />

conformational switching can be<br />

achieved in a different aromatic<br />

oligoamide system in which the<br />

protonation state of a repeating<br />

phenol functionality is controlled.<br />

[132]<br />

Oligoheterocyclic str<strong>and</strong>s consisting<br />

of alternating a,a’-connected<br />

pyridine <strong>and</strong> pyrimidine<br />

Figure 16. a) Chemical structure <strong>and</strong> folding of heptameric aromatic polyamide 34 (see<br />

Scheme 14 for representation of the hydrogen-bonding arrangement responsible for the folding).<br />

[129] b) In concentrated solutions, 34 dimerizes to form double helices. The double-helix<br />

structure is dissymmetrical <strong>and</strong> interconversion between the two equivalent forms occurs through<br />

a relative corkscrew motion of the two monomers. Exchange with the monomeric species is slow,<br />

as dissociation of the dimer involves unwinding of the helix structure. The ribbon representations<br />

are reprinted with permission from Ref. [129a].<br />

subunits adopt coiled structures because of the preferred<br />

transoid conformation at each heterocycle linkage. [133] However,the<br />

binding of a metal cation stabilizes the cisoid<br />

arrangement of the terpyridine-like coordination sites. In the<br />

presence of Pb II ions,for example,helix 36 is converted into<br />

an extended linear arrangement (a “[5]-rack” complex,<br />

Scheme 15). Again this process results in an enormous<br />

change in the molecular length,from 7.5 to 38 Š,<strong>and</strong> can<br />

be carried out reversibly <strong>and</strong> repeatedly by employing<br />

crypt<strong>and</strong> 37 to sequester <strong>and</strong> release Pb II ions depending on<br />

the pH value of the environment. [134] By contrast, a,a’-linked<br />

pyridine units naturally form linear arrangements. Although<br />

oligomers of this type are challenging to prepare,hydrazones<br />

can be used as isosteric replacements for alternate pyridine<br />

units,thus allowing preparation of linear pyridine–hydrazone<br />

sequences. These systems undergo the reverse process to that<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

93


Reviews<br />

Scheme 14. pH-controlled helix formation for aromatic polyamide<br />

heptamer 35. The insets show the hydrogen-bonding arrangements<br />

responsible for the helical structures. [130]<br />

illustrated for 36,namely,they form linear chains in solution<br />

which can be reversibly coiled around divalent metal ions. [135]<br />

Oligoheterocyclic systems such as 36 can also form<br />

double-helix structures. Whereas single/double helix interconversion<br />

for the aromatic oligoamide systems in Figure 16<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

is controlled by the solvent,concentration,<strong>and</strong> temperature,<br />

pyrimidine–pyridine oligomers can be made to form double<br />

helices by complexation to metal ions. Addition of Ag I ions to<br />

helical compound 38,for example,results in the formation of<br />

double-helical dimers of formula [(Ag I ) 2(38) 2] 2+ ,with the two<br />

metal ions bound at opposite ends of the chain <strong>and</strong> extensive<br />

intermolecular p–p stacking between the intertwined str<strong>and</strong>s<br />

(Figure 17). [136] The single/double helix interconversion process<br />

can be achieved reversibly in a manner analogous to that<br />

in Scheme 15.<br />

Figure 17. pH-switched contraction <strong>and</strong> extension by metal ion triggered<br />

single/double helix interconversion. [136] On models of 38 <strong>and</strong><br />

[(Ag I ) 2(38) 2] 2+ , SPr substituents have been omitted for clarity. The<br />

model structures are reprinted with permission from Ref. [136].<br />

Scheme 15. pH-switched helix formation using a metal-ion template. [134] Counterions (CF 3SO 3 ) <strong>and</strong> solvent (CH 3CN), which complete the Pb II<br />

coordination sphere in the [5]-rack complex [(Pb II ) 5(36)] 10+ , are omitted for clarity, as have the SPr substituents in the space-filling models. The<br />

space-filling representations are reprinted with permission from Ref. [134].<br />

94 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

A similar mechanism applies to 39,in which the two<br />

anthracene groups are held in a parallel arrangement,thereby<br />

creating a molecular tweezerlike cavity in which electronpoor<br />

guests such as 2,4,7-trinitro-9-fluorenone (TNF) can be<br />

bound (Scheme 16a). [137] Addition of cuprous ions,however,<br />

results in a cisoid–cisoid arrangement of the triheterocyclic<br />

unit so as to provide bipyridyl-like lig<strong>and</strong> environments for<br />

the metal ions. This separates the tweezer arms <strong>and</strong> releases<br />

the bound guest. Conversely,in 40,the oligoheterocyclic<br />

portion is designed so that the electron-rich binding cavity is<br />

only set up on ion binding (Zn II<br />

is used in this case,<br />

Scheme 16b).<br />

Salen/salophen lig<strong>and</strong> 41 is not intrinsically helical,but it<br />

forms helical structures on coordination to tetradentate metal<br />

ions such as Cu II or Ni II . [138] Electrochemical reduction of the<br />

Cu II -templated helix in a coordinating environment causes a<br />

reversible disruption of the helical secondary structure as the<br />

Cu I center prefers to adopt a pentacoordinate geometry in the<br />

absence of any tetrahedral lig<strong>and</strong> (Scheme 17).<br />

Finally,significant stimuli-induced conformational<br />

changes have also been achieved in a number of truly<br />

polymeric systems. More often than not,these are accompanied<br />

by detectable macroscopic changes in the material <strong>and</strong><br />

will be discussed further in Section 8.3.3.<br />

Scheme 16. Cation-induced substrate release <strong>and</strong> binding as a result of large-amplitude conformational changes. [137] a) On binding cuprous ions,<br />

U-shaped tweezerlike receptor 39 is forced to adopt a cisoid conformation at the pyridyl–pyrimidine linkages, thus separating the electron-rich<br />

tweezer “arms” <strong>and</strong> releasing the electron-poor guest. Dicuprous W-shaped derivative [(Cu I ) 2(39)] 2+ is only formed on addition of large excesses<br />

of [Cu(MeCN) 4] + (> 8 equiv). b) On binding zinc ions, W-shaped terpyridine 40 adopts a U shape, thus aligning the electron-rich anthracene<br />

moieties so that electron-poor guests can be intercalated in the cavity.<br />

Scheme 17. Formation of a helical complex ([Cu II (41)]) from a nonhelical lig<strong>and</strong> <strong>and</strong> electrochemically induced reversible unfolding of the helix. [138]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

95


Reviews<br />

2.1.5. Torsional Control Arising from <strong>Molecular</strong> Orbital Symmetry<br />

The stereochemical outcome of many chemical reactions<br />

arises from the relative movements of groups,which are<br />

determined by orbital symmetry considerations,most<br />

famously those governed by the Woodward–Hoffmann<br />

rules. [139] A full discussion of this area is beyond the scope<br />

of the current Review,but it is worth noting that such<br />

mechanisms for controlling submolecular motions in molecular-machine<br />

systems have been particularly underexploited<br />

to date. However,a purely electronically driven correlated<br />

rotational process has been observed in carbonium ion<br />

tricobalt complexes,driven by maintaining favorable overlaps<br />

between a molecular orbital centered on the three metal<br />

centers <strong>and</strong> an orbital belonging to the organic fragment as it<br />

moves between three degenerate conformational positions.<br />

[140] Recently,an interesting molecular dynamics study<br />

on 5-methyltropolone (42,Scheme 18),showed that proton<br />

transfer at the a-hydroxyketone should directly result in<br />

rotation of the methyl group at the 5-position. [141] This is<br />

explained by a coupling of the p system to the methyl group<br />

through hyperconjugation (involving two of the three methyl<br />

hydrogen 1s orbitals). Overlap of the HOMO of the methyl<br />

group with the LUMO of the ring is therefore strongest when<br />

the two out-of-plane hydrogen atoms straddle a single bond in<br />

the ring. As proton transfer results in tautomerization <strong>and</strong><br />

bond alternation round the<br />

ring,this flips the preferred<br />

orientation of the methyl<br />

group <strong>and</strong> a 608 rotation<br />

occurs.<br />

2.2. Controlling<br />

Configurational Changes<br />

Changes in configuration,<br />

[142]<br />

in particular cis–<br />

trans isomerization of double<br />

bonds,have been widely studied<br />

from theoretical,chemical,<strong>and</strong><br />

biological perspectives.<br />

[107,143] Although the<br />

small amplitude motion<br />

involved is not generally sufficient<br />

for direct exploitation<br />

as a machine,it can provide<br />

an extremely useful photoswitchable<br />

control mechanism<br />

for more complex systems<br />

(see below) <strong>and</strong>,in some<br />

cases,can even be harnessed<br />

to perform a significant<br />

mechanical task. Such systems<br />

represent some of the<br />

first examples of molecularlevel<br />

motion which can be<br />

controlled by the application<br />

of an external stimulus.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Scheme 18. Coupling of proton transfer to rotation of a methyl group<br />

in 5-methyltropolone (42). [141]<br />

As a prototypical example of this type of system,the<br />

photoisomerization of stilbenes has been extensively studied<br />

for well over 50 years. [144] As organic photochemistry developed,other<br />

important photochromic systems were discovered<br />

(Scheme 19): [145] the photoisomerization of azobenzenes, [146]<br />

the reversible electrocyclization of diarylethenes, [147]<br />

the<br />

photochromic reactions of fulgides, [148] the interconversion<br />

of spiropyrans with merocyanine (<strong>and</strong> the associated spirooxazine/merocyanine<br />

system), [149] as well as chiroptical switching<br />

in overcrowded alkenes (for example, 45,Scheme 21). [150]<br />

All these processes are accompanied by marked changes in<br />

both physical <strong>and</strong> chemical properties,such as color,charge,<br />

<strong>and</strong> stereochemistry,<strong>and</strong> a wide range of applications based<br />

on their switching properties have both been proposed <strong>and</strong>,in<br />

some cases,realized,from liquid crystal display technology to<br />

Scheme 19. Archetypical examples of photochromic systems that give a mechanical response. a), b) The<br />

cis–trans isomerization of stilbenes [144] <strong>and</strong> azobenzenes, [146] respectively. c) The reversible photochemical<br />

electrocyclization of a diarylethene. [147] This can be a competing process in the cis–trans isomerization of<br />

stilbenes (a) <strong>and</strong> is observed for a number of heterocyclic <strong>and</strong> non-heterocyclic aromatic compounds joined<br />

through a C=C bond. The methyl groups shown prevent irreversible loss of H 2 to give a tricyclic aromatic<br />

compound, while substitution of the double bond prevents cis!trans isomerization. d) The reversible<br />

photochemical electrocyclization of a fulgide. [148] Again, a number of substitution patterns are tolerated<br />

(cis–trans isomerization around either double bond may occur, but is undesirable). e) The photochemical<br />

interconversion of spiropyrans with their ring-open, zwitterionic merocyanine forms. [149] Analogues in which<br />

the benzylic carbon atom of the double bond is replaced with a nitrogen atom (spirooxazines) undergo a<br />

similar transition.<br />

96 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

variable-tint spectacles <strong>and</strong> optical recording media. [151] Most<br />

of these applications simply rely on the intrinsic electronic<br />

<strong>and</strong> spectroscopic changes on interconversion between the<br />

two species. There are some cases,however,in which small<br />

configurational changes can be harnessed in a more mechanical<br />

fashion <strong>and</strong> the resulting devices can be considered to be<br />

like molecular machines.<br />

Various configurational changes (in particular,photoisomerization<br />

of azobenzene) have been employed to alter<br />

peptide structures,thereby creating semisynthetic biomaterials<br />

whose activity can be controlled photonically. [97n–u] These<br />

systems,which amplify the small configurational change of<br />

the synthetic unit to cause a significant change in the peptide<br />

secondary structure,mimic to some extent the role of proline<br />

<strong>and</strong> peptidylprolyl cis–trans isomerases [152]<br />

in controlling<br />

protein activity in cells. Recently,it has been shown that the<br />

effective helix length can be controlled by photoisomerization<br />

in synthetic oligomers similar to those discussed in Section<br />

2.1.4 but composed of alternating 2,6-dicarboxamide <strong>and</strong><br />

m-(phenylazo)azobenzene units. [153] Incorporation of azobenzene<br />

or stilbene chromophores into dendritic molecules can<br />

result in interesting photochemical effects,as well as allowing<br />

amplification of the configurational change into large amplitude<br />

changes in molecular conformation or other properties.<br />

[154] These strategies are somewhat similar to the use of<br />

azobenzene units as phototriggers in polymeric synthetic<br />

materials (see Section 8.3.3).<br />

In small-molecule systems,converting a configurational<br />

change around a double bond into significant mechanical<br />

motion requires considerable ingenuity. Combining the<br />

reversible photoisomerization of an azobenzene with the<br />

“molecular-bearing” attributes of metallocene complexes<br />

(see Section 2.1.3),Aida <strong>and</strong> co-workers have created a pair<br />

of “molecular scissors” 43 (Scheme 20a). [155] The optically<br />

triggered change in the double-bond configuration brings<br />

about an angular change of position about the ferrocene<br />

“pivot” (green) <strong>and</strong> results in an opening <strong>and</strong> closing of the<br />

phenyl “blades” (red). Reversible switching is possible over a<br />

number of cycles,with the bite angle between the blades<br />

Scheme 20. a) “<strong>Molecular</strong> scissors” 43, in which photoisomerization<br />

of an azobenzene is converted into a 498 rotary motion around a<br />

metallocene bearing. [155] b) “<strong>Molecular</strong> hinge” 44, in which concerted<br />

configurational change of the two azobenzene units results in a<br />

change of approximately 908 in the angle between the planar xanthene<br />

units. [157]<br />

altered from approximately 98 when closed to more than 588<br />

when open.<br />

The concerted action of two azobenzene configurational<br />

switches [156] has been used to alter the angle between two<br />

planar moieties in a so-called “molecular hinge” 44 (Scheme<br />

20b). [157] In trans,trans-44,the two xanthene units are virtually<br />

coplanar. Photoisomerization to the cis,cis isomer occurs<br />

upon irradiation at 366 nm,thus resulting in an angle of<br />

almost 908 between the two aromatic ring systems. The high<br />

energy of the strained trans,cis form renders cis,cis-44<br />

remarkably thermally stable,but reisomerization can be<br />

achieved by irradiation at 436 nm.<br />

While investigating the use of overcrowded alkenes as<br />

chiroptical molecular switches, [150] Feringa et al. created a<br />

molecular brake, 45,in which the speed of rotation around an<br />

arene–arene single bond can be varied by changing the alkene<br />

configuration (Scheme 21). [158] Counterintuitively from the<br />

Scheme 21. Counterintuitive variation of kinetic barrier to rotation<br />

around an aryl–aryl bond by isomerization of the double bond in an<br />

overcrowded alkene. [158] Values for the free energy of activation (DG ° )<br />

for the rotary motion in each configuration are shown.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

two-dimensional representations in Scheme 21,the rate of<br />

rotation around the indicated bond is actually faster in cis-45<br />

than in trans-45 (demonstrated by 1 H NMR spectroscopy)—<br />

the naphthalene unit is flexible enough to bend away from the<br />

phenyl rotator unit,while the methylene protons on the other<br />

side of the double bond are rigidly held in equatorial <strong>and</strong> axial<br />

positions <strong>and</strong> present a significant steric barrier to the rotator.<br />

Unfortunately,the photochemical interconversion between<br />

the cis <strong>and</strong> trans forms is not efficient for this molecule,<br />

although the system st<strong>and</strong>s as proof (see also rotaxanes 87–89,<br />

Scheme 51) that a change in configuration can alter not just<br />

the optical properties or bulk orientations,but can also<br />

control aspects of large amplitude submolecular motions.<br />

The chiral helicity in molecules such as 45 causes the<br />

photochemically induced trans–cis isomerization to occur<br />

unidirectionally according to the h<strong>and</strong>edness of the helix.<br />

Accordingly,this system already possesses some of the<br />

features necessary for a unidirectional rotor <strong>and</strong>,indeed,<br />

the incorporation of one further chiral element [159] allowed<br />

the realization of the first synthetic molecular rotor capable of<br />

achieving a full <strong>and</strong> repetitive 3608 unidirectional rotation<br />

(Figure 18 a). [160,161] Irradiation (l > 280 nm) of this extraordinary<br />

molecule,(3R,3’R)-(P,P)-trans-46,causes chiral helicitydirected<br />

clockwise rotation of the upper half relative to the<br />

lower portion (as drawn),at the same time switching the<br />

configuration of the double bond <strong>and</strong> inverting the helicity to<br />

give (3R,3’R)-(M,M)-cis-46. However,this form is not stable<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

97


Reviews<br />

Figure 18. Operational sequence (a) <strong>and</strong> potential energy profile (b) of<br />

the first light-driven unidirectional 3608 molecular rotary motor,<br />

(3R,3’R)-46. [160] Note, the labels “stable” <strong>and</strong> “unstable” in (a) refer to<br />

thermodynamic stability, <strong>and</strong> the reaction coordinate in (b) does not<br />

correspond directly to the angle of rotation around the central bond.<br />

Even at the steady state (> 280 nm <strong>and</strong> > 608C), net fluxes occur in<br />

each of the isomer interconversions that make up the reaction cycle.<br />

The energy profile is adapted with permission from Ref. [161b].<br />

at temperatures above 55 8C as the methyl substituents on<br />

the cyclohexyl ring are placed in unfavorable equatorial<br />

positions. At ambient temperatures,therefore,the system<br />

relaxes through a second,thermally activated helix inversion<br />

to give (3R,3’R)-(P,P)-cis-46,with the substituents at either<br />

end of the double bond continuing to rotate in the same<br />

direction with respect to each other. Irradiation of (3R,3’R)-<br />

(P,P)-cis-46 (l > 280 nm) results in a second photoisomerization<br />

<strong>and</strong> helix inversion to give (3R,3’R)-(M, M)-trans-46.<br />

Again,the methyl substituents are placed in an energetically<br />

unfavorable position by the photoisomerization reaction <strong>and</strong><br />

thermal relaxation (this time temperatures > 60 8C are<br />

required [162] ) completes the 3608 rotation of the olefin<br />

substituents,thereby regenerating the starting species<br />

(3R,3’R)-(P,P)-trans-46. As each different step in the cycle<br />

involves a change in helicity,the directionality of the process<br />

can be monitored through the change in the circular dichroism<br />

spectrum at each stage. The four states can be differently<br />

populated depending on the precise choice of wavelength <strong>and</strong><br />

temperatures used,while irradiation at 280 nm at > 60 8C<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

results in continuous 3608 rotation. [160] Figure 18 b shows the<br />

overall potential-energy changes during a full 3608 rotation.<br />

In terms of mechanism,the process essentially involves<br />

switching between two asymmetric potential energy surfaces<br />

(one each for the cis <strong>and</strong> trans diastereomers) which are<br />

spatially separated (in terms of their position along the<br />

reaction coordinate for rotation around the central bond).<br />

Furthermore,this switching process is directional (see above)<br />

<strong>and</strong> so this system bears many of the hallmarks of a<br />

dichotomous,stepwise traveling-wave ratchet (see Section<br />

1.4.2.1). The potential-energy minima <strong>and</strong> maxima on<br />

the cis <strong>and</strong> trans surfaces are of course actually slightly<br />

different,<strong>and</strong> so elements of a flashing ratchet are also<br />

present—such hybrids are perfectly acceptable under Brownian<br />

ratchet theory. [39f]<br />

Note that if 46 is irradiated at wavelengths greater than<br />

280 nm <strong>and</strong> at temperatures above 608C the system will reach<br />

a steady state at which point the bulk distribution of<br />

diastereomers no longer changes. Crucially,however,this<br />

steady state is not an adiabatic equilibrium. The steady state is<br />

maintained by a process (Figure 18) that features nonzero<br />

fluxes (corresponding to directional rotation of one component<br />

with respect to the other) between various pairs of<br />

diastereomers. In other words,as long as an external source of<br />

light <strong>and</strong> heat is supplied,the steady state in Figure 18 is<br />

effectively maintained by a cyclic process A!B!C!D!A<br />

(see Section 1.4.1),with the arrow indicating a net flux. Thus,<br />

even at the steady state, 46,<strong>and</strong> other molecules of this type,<br />

behave as directional rotary motors.<br />

As the photoisomerization process in such systems is<br />

extremely fast (< 300 ps), [163] the rate-limiting step in the<br />

operation of 46 is the slowest of the thermally activated<br />

isomerization reactions. The effect of molecular structure on<br />

this rate has been investigated in a series of derivatives. An<br />

ethyl substituent at the two stereogenic centers results in a<br />

moderate increase in the rates of thermal relaxation,probably<br />

because of greater steric hindrance when these groups occupy<br />

the unfavorable equatorial positions. [164] The isopropyl-substituted<br />

homologue showed a further increase in the rate for<br />

the thermal relaxation step between the cis diastereomers<br />

(fast even at 60 8C),but this trend was not followed for the<br />

equivalent step between the two trans diastereomers,which<br />

was very significantly retarded. This situation allowed the<br />

isolation of an intermediate meso-like isomer of the form<br />

(P,M)-trans,thus demonstrating that the thermal relaxation<br />

step in fact comprises two consecutive helix inversions, [164] inline<br />

with an earlier theoretical prediction. [162b] A second<br />

generation of motors (47–49,Scheme 22) was created in<br />

which “rotator” <strong>and</strong> “stator” sections of the molecule are<br />

differentiated (with a view to future applications; see<br />

Section 8.1) <strong>and</strong> in which only one stereogenic center is<br />

sufficient to direct the rotation. [165] For all the second<br />

generation motors shown in Scheme 22,the slowest thermal<br />

isomerization step has a lower kinetic barrier than in 46; in<br />

general,smaller bridging groups at Y <strong>and</strong>,in particular,at X<br />

result in the lowest activation barriers. Not all the kinetic<br />

effects of the structural changes turned out to be intuitive,<br />

however. Motor 48 contains an even smaller upper portion,<br />

yet the free energy of activation for the thermal isomerization<br />

98 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 22. General structure of second-generation light-driven<br />

unidirectional rotors 47, 48, <strong>and</strong> the fastest member of this series,<br />

49. [165] Boc = tert-butyloxycarbonyl.<br />

process in this molecule is actually higher than its phenanthrene<br />

analogue (that is, 47;X= CH 2,Y= CH=CH,R 1 = R 2 =<br />

H). [165c] The decreased steric hindrance in 48 lowers the<br />

ground-state energy of the “unstable” isomer more than it<br />

lowers the transition-state energy for the thermal isomerization<br />

process. Conversely,a derivative bearing donor <strong>and</strong><br />

acceptor substituents (47;X= S,Y = S,R 1 = NMe 2,R 2 = NO 2)<br />

exhibited thermal relaxation rates more comparable with less<br />

sterically encumbered examples. [165d] Furthermore,this molecule<br />

can operate using visible light,while protonation of the<br />

amine substituent led to slightly improved ratios for the<br />

photostationary state,which were reached in shorter irradiation<br />

times,but were then followed by slower thermal<br />

relaxation steps. It appears,therefore,that electronic effects<br />

also play an important role in the mechanism of these systems.<br />

This was further emphasized by derivative 49,in which an<br />

amine in the upper half of the molecule is directly conjugated<br />

with a ketone in the lower half. The consequent increase in<br />

single-bond character of the central olefin results in greatly<br />

increased rates for the thermal isomerization steps. [165e]<br />

Fast rotation rates have also been achieved with cyclopentane<br />

analogue 50 which can be accessed in higher<br />

synthetic yields than the six-membered<br />

ring compounds (Scheme 23). [166]<br />

Despite the greater conformational<br />

flexibility of the five-membered ring,a<br />

significant difference in the energies<br />

between the pseudoequatorial <strong>and</strong><br />

pseudoaxial positions of the appended<br />

Scheme 23. Thirdgenerationunidirectional<br />

rotor 50. [166] In<br />

this case “ax” refers<br />

to the pseudoaxial<br />

orientation for substituents<br />

on the<br />

cyclopentyl ring.<br />

methyl groups still exists,<strong>and</strong> unidirectional<br />

rotation occurs. Both 49 <strong>and</strong> 50<br />

suffer from lower photostationary state<br />

ratios, [167] while the thermal relaxation<br />

steps for 49 are accompanied by a small<br />

amount of thermal back-isomerization.<br />

[165e,166a] However,neither of these<br />

factors affects the integrity of the process,as<br />

in all cases the thermal relaxa-<br />

tion steps are entirely unidirectional,thus ensuring the<br />

motion is ratcheted. Only the overall photon efficiency for<br />

rotation (a combination of quantum yield for the isomerization<br />

<strong>and</strong> the photostationary state ratio) is reduced.<br />

Recently,Feringa <strong>and</strong> co-workers have returned to the<br />

theme of using isomerization around the central olefin to<br />

control the rate of rotation of an aryl group in motor molecule<br />

51 (Scheme 24). [168] Although 51 displays more complex<br />

photochemistry than the previous motor molecules,it still<br />

exhibits unidirectional rotation around the olefinic bond. The<br />

Scheme 24. Rotational scheme for molecular motor 51 in which each<br />

diastereoisomer has a different barrier to rotation around the aryl–aryl<br />

bond indicated. [168] The terms “stable” <strong>and</strong> “unstable” refer to the<br />

thermodynamic stabilities of each isomer.<br />

rate of rotation around the aryl–aryl single bond was<br />

determined experimentally for each diastereoisomer <strong>and</strong><br />

found to follow the order cis stable > cis unstable > trans stable ><br />

trans unstable. Similar to switch 45 (Scheme 21),therefore,the<br />

sp 3 -hybridized side of the dihydronaphthothiopyran part<br />

exerts the greater steric hindrance on the aryl rotator,<br />

particularly when the methyl substituent is in the less-favored<br />

equatorial position. Of course,the unidirectionality of the<br />

conversions between the four isomers is immaterial to the<br />

switching of the rotation rates.<br />

3. Controlling Motion in Supramolecular Systems<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Using an external stimulus to modulate the binding<br />

affinity of a host for a guest is the simplest expression of<br />

controllable molecular recognition. [169] A wide variety of<br />

stimuli can be employed to bring about changes,not just in<br />

geometric configuration,but also in electronic arrangement<br />

<strong>and</strong> environmental influences to modulate noncovalent<br />

interactions. Switchable host–guest systems teach us much<br />

about the nature of noncovalent interactions <strong>and</strong> how to<br />

manipulate them,although the requirement for kinetic<br />

association of components in a molecular machine rules out<br />

simple host–guest complexation where binding does not bring<br />

about a change in the conformation of either species or where<br />

transport of the guest between sites within the host is slow<br />

with respect to exchange with the bulk. For example,myosin<br />

must move along a track to which it is kinetically associated<br />

through a series of sequential binding events to bring about<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

99


Reviews<br />

muscle contraction; no mechanical task occurs through simple<br />

“on”/“off” binding to the track by myosin molecules from the<br />

bulk. [170] In other words,simple host–guest/supramolecular<br />

systems cannot function as nanoscale mechanical machines<br />

unless restrictions on the motion of the unbound species apply<br />

or the binding event causes a mechanical (that is,conformational)<br />

change in one of the molecular components. A stimuliinduced<br />

molecular recognition event is neither a sufficient nor<br />

necessary condition for the construction of a molecular-level<br />

machine.<br />

3.1. Switchable Host–Guest Systems<br />

Whilst the requirement for kinetic association during the<br />

operation of the machine means that many switchable host–<br />

guest systems do not have the potential to act as mechanical<br />

machines,many still do. We have already seen that certain<br />

synthetic allosteric systems transmit binding at one site to a<br />

remote receptor through binding-induced conformational<br />

motions (see Sections 2.1.3 <strong>and</strong> 2.1.4). In other conformation-switching<br />

molecular machines,intermolecular binding<br />

events are both the stimulus <strong>and</strong> outcome of the molecularlevel<br />

motion. An early example of a switchable host–guest<br />

system which demonstrates molecular-machine-like characteristics<br />

is (E)/(Z)-52 (Scheme 25 a) developed by Shinkai<br />

et al. [171] These photoresponsive “molecular tweezers” exhibit<br />

high selectivity for binding large cations such as Rb + when in<br />

the cis form <strong>and</strong> selectivity for small cations such as Na + in the<br />

trans form. Furthermore,the presence of Rb + ions in solution<br />

increases the proportion of (Z)-52 at the photostationary state<br />

<strong>and</strong> decreases the rate of the thermal Z!E transformation.<br />

Subsequently,the same research group developed the “tailbiting”<br />

crown ether switchable receptor (E)/(Z)-53·H +<br />

(Scheme 25b). [172] In the (Z)-53·H + form,intramolecular<br />

binding of the ammonium ion to the crown ether prevents<br />

recognition of other cations in the medium,while thermal<br />

isomerization to give (E)-53·H + restores the properties of the<br />

crown ether necessary for binding alkali metal cations. It is<br />

also observed that the rate of Z!E isomerization is lowered<br />

(by 1.6–2.2 fold) for (Z)-53·H + relative to the deprotonated<br />

control compound (Z)-53. Crucially,the rate of this reaction<br />

increases as the concentration of K + ions in solution increases.<br />

Both these systems can therefore be viewed as primitive<br />

molecular machines in which a binding<br />

event at the crown ether unit(s)<br />

affects the isomerization processes of<br />

the azobenzene moiety with which it is<br />

kinetically associated.<br />

3.2. Intramachine Ion Translocation<br />

Metal–lig<strong>and</strong> binding interactions<br />

are often relatively kinetically inert<br />

<strong>and</strong> careful structural design has produced<br />

systems in which intramolecular<br />

motions are significantly favored<br />

over intermolecular exchange. [173]<br />

Scheme 25. Two of the earliest molecular machines. a) Photoresponsive<br />

molecular tweezers (E)/(Z)-52. [171] b) “Tail-biting” crown ether<br />

(E)/(Z)-53. [172]<br />

System 54 4+ (Scheme 26) consists of a coordinatively unsaturated<br />

Cu II center covalently linked to a redox-active Ni II –<br />

cyclam unit. [174] In this form,chloride anions bind strongly to<br />

the copper center,filling the vacant coordination site<br />

([54(Cl)] 3+ ). Electrochemical oxidation of the nickel center<br />

to Ni III dramatically increases its affinity for anions so that the<br />

chloride translocates to this new,more energetically favorable<br />

site. The motion is completely reversible on reduction of the<br />

nickel ion. In principle,the switching of the anion position<br />

could be an intermolecular process involving either free<br />

anions in solution or more than one molecule of 54. However,<br />

the anion translocation was demonstrated experimentally not<br />

to be concentration-dependent,in contrast to an analogous<br />

three-component system (Cu receptor + Ni receptor + Cl )<br />

which exhibited strongly concentration-dependent behavior.<br />

Thermodynamic comparisons of the two systems suggest that<br />

the dominant mechanism in 54 is an intramolecular translocation,brought<br />

about by folding of the ditopic receptor. [175]<br />

Scheme 26. Redox-driven intramolecular anion translocation. [174]<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

100 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Calix[4]arene 55 reversibly translocates metal cations<br />

through protonation of the tertiary nitrogen atom<br />

(Scheme 27). [176] Dynamic 1 H NMR experiments indicate a<br />

high kinetic barrier to dissociation of the metal ion from the<br />

Scheme 27. Translocation of a silver cation between two sites in a<br />

calix[4]arene cavity: a “molecular syringe”. [176]<br />

neutral receptor,thus suggesting that upon protonation,<br />

translocation occurs intramolecularly through the cavity<br />

defined by the aromatic rings in a manner reminiscent of<br />

the action of a syringe. However,the protonated form exhibits<br />

much faster cation exchange with the bulk so the integrity of<br />

the stimuli-induced return stroke is probably less well<br />

maintained. [176]<br />

The first redox-driven cation-translocation system was<br />

reported by Shanzer <strong>and</strong> co-workers in helical complex<br />

[Fe III ·56],which exhibits strong evidence for a fully intramolecular<br />

process (Scheme 28). [177] The system exploits the<br />

preference of Fe III <strong>and</strong> Fe II ions for hard <strong>and</strong> soft lig<strong>and</strong>s,<br />

respectively. Reduction of [Fe III (56)] with ascorbic acid<br />

Scheme 28. Redox-switched intramolecular cation translocation in a<br />

triple-str<strong>and</strong>ed helical complex. [177] Reductant: ascorbic acid; oxidant:<br />

(NH 4) 2S 2O 8.<br />

generates [Fe II (56)(H) 3] 2+ which has spectral properties<br />

characteristic of the [Fe II (bipyridyl) 3] coordination sphere.<br />

Subsequent reoxidation ((NH4) 2S2O8) returns the system to<br />

its original state. The equivalent intermolecular process<br />

between two monotopic lig<strong>and</strong>s does not occur under the<br />

same conditions,while even the translocation of the cation in<br />

56 is relatively kinetically slow,thus suggesting an intramolecular<br />

process. [178]<br />

Pseudorotaxanes—supramolecular complexes in which a<br />

macrocyclic host encapsulates a linear,threadlike guest—are<br />

subject to the same issues regarding kinetic stability as other<br />

host–guest complexes. [179–181] While these complexes are<br />

model systems for kinetically locked analogues (rotaxanes, [182]<br />

see Section 4),their dynamic properties are similar to other<br />

supramolecular systems. Many interesting pseudorotaxane<br />

devices have been created [183] with functions including:<br />

reversible formation <strong>and</strong> dethreading by using a number of<br />

stimuli; [122,123d,184]<br />

switching between different preferred<br />

guests; [185] shuttling between two binding sites within the<br />

same guest; [186] <strong>and</strong> control of intracomplex electron-transfer<br />

reactions. [187] However,only in the case of kinetically stable<br />

pseudorotaxanes—systems where the components do not<br />

exchange with the bulk during the operation of the machine—<br />

are there sufficient restrictions on the motions of the unbound<br />

species for them to feature controlled mechanical behavior<br />

<strong>and</strong> they are otherwise best considered as supramolecular<br />

devices not mechanical machines. [188]<br />

4. Controlling Motion in <strong>Mechanical</strong>ly Bonded<br />

<strong>Molecular</strong> Systems<br />

4.1. Basic Features<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Catenanes are chemical structures in which two or more<br />

macrocycles are interlocked,while in rotaxanes one or more<br />

macrocycles are mechanically prevented from dethreading<br />

from a linear unit by bulky “stoppers” (Figure 19). [189] Even<br />

though their components are not covalently connected,<br />

catenanes <strong>and</strong> rotaxanes are molecules (not supramolecular<br />

complexes) as covalent bonds must be broken to separate the<br />

constituent parts. In these structures, [190] the mechanical bond<br />

severely restricts the relative degrees of freedom of the<br />

Figure 19. Schematic representations of: a) a [2]catenane <strong>and</strong> b) a<br />

[2]rotaxane. [192] The arrows show possible large-amplitude modes of<br />

movement for one component relative to another. In a [2]rotaxane (b),<br />

these are defined as I: translation <strong>and</strong> II: pirouetting. In a [2]catenane,<br />

(a), however, motion I can be considered as pirouetting of the blue<br />

ring around the orange one or, equally, translation of the orange ring<br />

around the blue one.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

101


Reviews<br />

components in several directions,while often permitting<br />

motion of extraordinarily large amplitude in an allowed<br />

vector. This situation is in many ways analogous to the<br />

restriction of movement imposed on biological motors by a<br />

track [170] <strong>and</strong> is one reason interlocked structures continue to<br />

play a central role in the development of synthetic molecular<br />

machines. [191]<br />

The large-amplitude submolecular motions particular to<br />

catenanes <strong>and</strong> rotaxanes can be divided into two classes<br />

(Figure 19): pirouetting of the macrocycle around the thread<br />

(rotaxanes) or the other ring (catenanes) <strong>and</strong> translation of<br />

the macrocycle along the thread (rotaxanes) or around the<br />

other ring (catenanes). By analogy to the stereochemical term<br />

“conformation”,which refers to geometries that can formally<br />

be interconverted by rotating about<br />

covalent bonds,the relative positioning<br />

of the components in interlocked<br />

molecules (<strong>and</strong> supramolecular complexes)<br />

is often referred to as a “coconformation”.<br />

[142]<br />

For a long time,synthetic<br />

approaches to mechanically interlocked<br />

structures relied on statistical<br />

or lengthy covalent-directed<br />

approaches. [193] The development of<br />

supramolecular chemistry,however,<br />

allowed chemists to apply noncovalent<br />

interactions to synthesis,thus<br />

resulting in many template methods<br />

(“clipping” <strong>and</strong> “threading”) [194]<br />

to<br />

catenanes <strong>and</strong> rotaxanes being developed.<br />

[195] In such syntheses,noncovalent<br />

binding interactions between the<br />

components often “live-on” in the<br />

interlocked products. These can ultimately<br />

be manipulated to effect positional<br />

displacements of one component<br />

with respect to another. Much<br />

attention has been given to the submolecular<br />

motions within these structures<br />

at equilibrium. We shall briefly<br />

cover the inherent large amplitude<br />

motion in rotaxanes as it is instructive<br />

for the consideration of stimuliinduced<br />

control of motion in these<br />

structures (see Sections 4.3–4.6).<br />

4.2. Shuttling in Rotaxanes: Inherent<br />

Dynamics<br />

Shuttling is the movement of a macrocycle along the<br />

linear thread component of a rotaxane. This motion takes the<br />

form of a r<strong>and</strong>om walk powered by Brownian motion that is<br />

constrained to one dimension by the thread <strong>and</strong> to translational<br />

displacement boundaries by the bulky stoppers. By<br />

virtue of the template methods employed in interlocked<br />

molecule synthesis, [195] rotaxanes without sites of attractive<br />

interaction between the macrocycle <strong>and</strong> thread are relatively<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

rare. [196] It is more common that the thread consists of one or<br />

more recognition elements,or “stations”,for the macrocycle(s);<br />

shuttling therefore becomes the movement on,off,<br />

<strong>and</strong> between such stations,with rates dependent on the<br />

strength of the intercomponent interactions.<br />

4.2.1. Observation of Shuttling in Degenerate, Two Binding Site<br />

<strong>Molecular</strong> Shuttles<br />

Stoddart <strong>and</strong> co-workers demonstrated that 57 4+ ,the first<br />

[2]rotaxane to be constructed with two well-defined noncovalent<br />

binding sites,exhibits shuttling behavior (Scheme<br />

29a). [197] 1 H NMR experiments were consistent with the<br />

macrocycle moving between the two hydroquinol stations in<br />

Scheme 29. a) The first “molecular shuttle” 57 4+ . [197] b) Peptide-based degenerate molecular<br />

shuttles 58–61. [199]<br />

a temperature-dependent fashion. Directly analogous situations<br />

have subsequently been demonstrated for many rotaxanes<br />

bearing the same,or similar,macrocycles <strong>and</strong> multiple<br />

electron-rich aromatic rings on the thread. [198]<br />

Similar effects are seen with other multistation rotaxane<br />

systems. Shuttling dynamics have been systematically studied<br />

for a series of peptide-based molecular shuttles in which two<br />

glycylglycine stations are separated by aliphatic linkers. [199] In<br />

CDCl 3,the macrocycle in 58–60 shuttles between the two<br />

degenerate peptide stations rapidly at room temperature<br />

(Scheme 29b). The shuttling mechanism—a simplified profile<br />

is shown in Figure 20—requires at least partial rupture of the<br />

intercomponent interactions at one station before formation<br />

102 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 20. Idealized free-energy profile for movement<br />

between two identical stations in a degenerate<br />

molecular shuttle. The height of the barrier DG °<br />

contains two components: the energy required to<br />

break the noncovalent interactions holding it to the<br />

station <strong>and</strong> a distance-dependent diffusional component.<br />

of new interactions at a second station. If the<br />

kinetic barrier can be made to be significantly<br />

larger than the available thermal energy,the<br />

shuttling will stop. This can be achieved for 60<br />

by introduction of a bulky N-tosyl (NTs)<br />

moiety (giving 61,Scheme 29b). Shuttling is<br />

restored upon removal of the N-tosyl barrier.<br />

In single binding site rotaxanes,hydrogenbond-disrupting<br />

solvents such as [D 6]DMSO<br />

break up the interactions between the benzylic<br />

amide macrocycle <strong>and</strong> peptide units, [200]<br />

<strong>and</strong> solvent composition indeed has a profound<br />

effect on the shuttling rate in 58–60.As<br />

little as 5% [D 4]methanol increases the<br />

shuttling rate by more than two orders of<br />

magnitude in halogenated solvents. [199,201] In a<br />

different hydrogen-bonded system,the shuttling<br />

rate could be controlled by deprotonation<br />

of a phenol moiety located between two<br />

degenerate stations. [202] Noncovalent interaction<br />

between the phenolate <strong>and</strong> counterion<br />

was found to significantly slow—but not<br />

prevent—shuttling,while the precise rate<br />

could again be fine-tuned by variation of the<br />

solvent composition.<br />

The reversible introduction of large<br />

kinetic barriers to shuttling has also been<br />

achieved without formation of covalent<br />

bonds. [203] In rotaxane 62,the crown ether<br />

normally shuttles between the two bipyridinium<br />

stations (Scheme 30). The introduction<br />

of Cu I ions,however,forms a kinetically stable dimeric<br />

complex,thus blocking movement of the ring. An ionexchange<br />

resin can be used to sequester the metal <strong>and</strong> restore<br />

shuttling. [204]<br />

4.2.2. A Physical Model of Degenerate, Two Binding Site<br />

<strong>Molecular</strong> Shuttles<br />

Angew<strong>and</strong>te<br />

Chemie<br />

An interesting structural effect is observed on increasing<br />

the length of the spacer in degenerate shuttles. Although<br />

ostensibly not involved in any interactions with the macrocycle,extension<br />

of the alkyl chain from 58 to 59 (Scheme 29b)<br />

reduces the rate of shuttling by a factor corresponding to an<br />

increase in activation energy of 1.2 kcalmol 1 —an effect<br />

solely of the increased distance the macrocycle must<br />

Scheme 30. Switchable degenerate shuttling through reversible formation of kinetically stable<br />

dimeric complexes. [204a]<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

103


Reviews<br />

travel. [199] This effect is most clearly understood by considering<br />

the macrocycle as a particle moving along a one-dimensional<br />

potential energy (rather than free-energy) surface<br />

(Figure 21). At any point on the potential-energy surface,the<br />

gradient of the line gives the force exerted on the macrocycle<br />

by the thread. When the macrocycle is in the vicinity of a<br />

station,hydrogen bonds <strong>and</strong>/or other attractive noncovalent<br />

interactions exert large forces (typically varying as a high<br />

power of the inverse distance r n ) on the macrocycle,<br />

opposing its motion. When the macrocycle is on the thread<br />

between the stations,the hydrogen bonds <strong>and</strong> other noncovalent<br />

binding interactions are broken <strong>and</strong> no forces are<br />

exerted on the macrocycle by the thread (that is,the gradient<br />

of the line is zero). The rate at which the macrocycles escape<br />

from each station is given by a st<strong>and</strong>ard Arrhenius equation,<br />

which depends on the depth of the energy well <strong>and</strong> the<br />

temperature. However,this is not the only factor involved in<br />

determining the rate at which macrocycles move between the<br />

two stations; there must also be some distance-dependent<br />

diffusional factor in the function describing rate of shuttling<br />

(Figure 21).<br />

The experimentally determined values of DG ° for 58 <strong>and</strong><br />

59 have also been reconciled with a quantum-mechanical<br />

description of the shuttling process. [205] Calculation of the<br />

wavefunctions for the system shows that an increase in<br />

distance between the stations does not,of course,change the<br />

activation energy for cleavage of hydrogen bonds but rather<br />

the effect is to widen the free-energy potential well<br />

(Figure 20). As the well widens,it possesses a higher density<br />

of states per unit energy (just like the simple “particle-in-abox”<br />

model). The closer the levels are to one another,the<br />

more readily thermally populated they are or,in other words,<br />

the larger is the partition function <strong>and</strong> therefore also the<br />

DG ° value.<br />

Figure 21. Idealized potential energy of the macrocycle in a two-station, degenerate molecular shuttle. The<br />

potential-energy surface shows the effect of the interaction between the macrocycle <strong>and</strong> thread on the<br />

energy of the macrocycle (ignoring any complicating factors such as folding). Chemical potential energies<br />

(DE values) generally follow similar trends to free energies (DG values, Figure 20) but there are some<br />

important differences; for example, the activation free energy of shuttling DG ° corresponds to the energy<br />

required for the macrocycle to move all the way to the new binding site (that is, it includes a contribution<br />

for the distance the ring has to move along the track to reach the other station) whereas DE ° (shown here)<br />

represents the energy required for the macrocycle to escape the forces exerted through noncovalent binding<br />

interactions at a station. The main plot shows the DE value in terms of the position of the macrocycle along<br />

the vector of the thread; the minor plot shows the DE value in terms of the position of the macrocycle<br />

orthogonal to the vector of the thread, thus illustrating that the thread genuinely behaves as a onedimensional<br />

potential-energy surface for the macrocycle.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

The nature of the wavefunctions at energies close to the<br />

top of the barrier is intriguing,as under this energy regime the<br />

maximum probability of finding the macrocycle is over the<br />

aliphatic spacer. The shuttling process can therefore be<br />

thought of in terms of a function of free energy (Figure 20)<br />

as long as we remember that the height of the DG ° barrier is<br />

affected by both binding strength <strong>and</strong> distance between the<br />

stations. The behavior of the macrocycle is similar to a cart<br />

moving along a roller coaster track shaped like the double<br />

potential minima in Figure 20. At low temperatures,the cart<br />

mostly resides on the stations (oscillating with small amplitude<br />

in the troughs). As energy approaches the value of the<br />

barrier,the cart spends most of its time passing over the<br />

barrier (occupying co-conformations close to the transition<br />

state for movement between the stations). At higher energies<br />

still,the cart is most likely to be found at the extremes of its<br />

translational motion (once again occupying ground-state coconformations<br />

at the stations).<br />

4.3. Stimuli-Responsive <strong>Molecular</strong> Shuttles: Translational<br />

<strong>Molecular</strong> Switches<br />

As we have seen,the rate at which the macrocycle moves<br />

between the stations by Brownian motion in molecular<br />

shuttles can be regulated by temperature <strong>and</strong>,in some<br />

cases,solvent composition or binding events. However,the<br />

principle of detailed balance (Section 1.4.1) tells us that no<br />

useful task can be performed by such a process,even if the two<br />

stations are different. A more important kind of control is one<br />

in which the net location of the macrocycle is changed in<br />

response to an applied stimulus. This breaks detailed balance<br />

<strong>and</strong>,as the system relaxes back to equilibrium,the biased<br />

Brownian motion of the macrocycle can be used to perform a<br />

mechanical task. [206]<br />

4.3.1. Single Binding Site, Stimuli-<br />

Responsive <strong>Molecular</strong> Shuttles<br />

A limited amount of control over<br />

the position of the macrocycle can be<br />

introduced into single recognition site<br />

rotaxanes if the binding affinity can be<br />

affected by a stimulus. An important<br />

first step towards photoswitchable<br />

molecular shuttles was [2]rotaxane<br />

63 4+ (Scheme 31). [207] This rotaxane is<br />

closely related to 57 4+ ,but in 63 4+<br />

there is only one electron-rich station<br />

for the macrocycle <strong>and</strong> the bulky<br />

stoppers are redox-active ferrocene<br />

groups. A laser flash photolysis pulse<br />

within the charge transfer b<strong>and</strong> of<br />

63 4+ induces electron transfer from<br />

the dioxyarene to the cyclophane,<br />

thereby generating an intimate radical<br />

ion pair. Charge recombination is<br />

normally fast compared with any<br />

competing processes such as solvent<br />

104 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 31. Conformational dynamics <strong>and</strong> electron-transfer processes possible in [2]rotaxane 63 4+ : a “single-station” molecular shuttle. [207]<br />

penetration or spatial separation of the radical ions. The<br />

flexible thread,however,allows close contact of the electronrich<br />

ferrocenyl stoppers with the cyclophane through a<br />

secondary p-stacking interaction (also observed in the solid<br />

state). This electronic coupling allows some of the radical ion<br />

pairs (ca. 25%) to undergo a secondary electron-transfer step<br />

in which the electron hole is transferred to one of the<br />

ferrocenyl stoppers. The interactions between the cyclophane<br />

<strong>and</strong> the stopper are therefore “switched off” <strong>and</strong> the molecule<br />

unravels. This spatially remote charge-separated state is<br />

relatively long-lived,so that some shuttling of the cyclophane<br />

away from the oxidized stopper may occur,driven by<br />

electrostatic repulsion. However,there is no direct evidence<br />

for this transient shuttling,or means to control the position of<br />

the cyclophane in the spatially remote charge-separated state.<br />

The environment-sensitive nature of the intercomponent<br />

hydrogen bonds in peptidic [2]rotaxanes allows control over<br />

effects arising from communication between the thread <strong>and</strong><br />

macrocycle (Scheme 32a). Unlike the free sarcoglycine<br />

thread,rotaxane 64 exhibits only one (E) tertiary amide<br />

rotamer in apolar solvents,as the Z rotamer allows the<br />

formation of fewer favorable intercomponent hydrogen<br />

bonds (Scheme 32a). In hydrogen-bond-competing solvents,<br />

however,the interactions between the thread <strong>and</strong> macrocycle<br />

are “switched off” <strong>and</strong> both rotamers are observed. [208] It was<br />

found that a similar effect could switch on <strong>and</strong> off an induced<br />

circular dichroism (ICD) response between the intrinsically<br />

achiral benzylic amide macrocycle <strong>and</strong> the chiral center in the<br />

thread in chiral dipeptide [2]rotaxane 65 (Scheme 32b). [209]<br />

In two related systems,it has been possible to control the<br />

position of a cyclodextrin (CD) macrocycle on threads<br />

containing azobenzene <strong>and</strong> stilbene units (Scheme 33). [210,211]<br />

The cyclodextrin spends most of its time over the central<br />

aromatic units in the E isomers of [2]rotaxanes 66 <strong>and</strong> 67 in<br />

aqueous media. Irradiation at suitable wavelengths results in<br />

photoisomerization of the N=N <strong>and</strong> C=C bonds in 66 <strong>and</strong> 67,<br />

respectively,to the Z forms. The steric dem<strong>and</strong>s of the<br />

“kinked” thread require the cyclodextrin to be positioned<br />

away from the central unit. Even in the E isomers,however,<br />

the cyclodextrin does not sit exclusively over the central unit.<br />

This is in fact crucial for allowing photoisomerization to<br />

occur—an analogue of 66 which lacks the ethylene spacer<br />

(light blue) between the azobenzene <strong>and</strong> viologen units<br />

undergoes no photoisomerization as the trans chromophore is<br />

fully encapsulated by the ring. [210b,212] Interestingly,in 67,<br />

despite having a symmetrical thread,the displacement has<br />

been shown to be unidirectional,with the narrower 6-rim of<br />

the cyclodextrin always closest to the Z olefin. [211,213] It has<br />

been noted in a variety of studies that the asymmetry of<br />

cyclodextrins can result in the selective formation of orientational<br />

isomers when constructing pseudorotaxane or rotaxane<br />

architectures. [214] Although the observation of such isomers<br />

requires nonsymmetrical thread portions,an asymmetric<br />

substrate,such as a cyclodextrin ring,should itself satisfy<br />

the requirement of symmetry breaking for the creation of a<br />

full ratchet mechanism even on a symmetric track (see<br />

Section 1.4.2).<br />

4.3.2. A Physical Model of Two Binding Site, Stimuli-Responsive<br />

<strong>Molecular</strong> Shuttles<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Rotaxanes in which the macrocycle can be translocated<br />

between two or more well-separated stations in response to<br />

an external signal should,in principle,provide a greater level<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

105


Reviews<br />

Scheme 32. Environment-sensitive communication between interlocked components in single-station dipeptide [2]rotaxanes. a) In C 2D 2Cl 4,<br />

hydrogen bonding between the thread <strong>and</strong> macrocycle stabilizes the E rotamer so that this is the sole conformation observed by NMR<br />

spectroscopy. In [D 6]DMSO, the hydrogen bonding is switched off <strong>and</strong> both rotamers are observed, as they are for the free thread in all<br />

solvents. [208] b) In CHCl 3, the achiral macrocycle is held close to the chiral center, thereby conferring a chiral nature on its conformation which, in<br />

turn, affects the conformation of the C-terminal diphenylmethyl moiety, thus resulting in an ICD response. In MeOH, the hydrogen bonding is<br />

switched off <strong>and</strong> the communication between the components is lost. [209]<br />

Scheme 33. Photoresponsive single-station shuttles 66 [210] <strong>and</strong> 67, [211] based on azobenzene <strong>and</strong><br />

stilbene units, respectively.<br />

of positional control. As seen in Section 4.2.2,in any rotaxane<br />

the macrocycle distributes itself between the available binding<br />

sites according to the difference in the macrocycle binding<br />

energies <strong>and</strong> the temperature. If a suitably large difference in<br />

macrocycle affinity exists between two stations,the macrocycle<br />

resides overwhelmingly in one positional isomer or coconformation.<br />

[142] In stimuli-responsive molecular shuttles,an<br />

external trigger is used to chemically modify the system <strong>and</strong><br />

alter the noncovalent intercomponent interactions such that<br />

the second macrocycle binding site becomes energetically<br />

more favored,thus causing translocation of the macrocycle<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

along the thread to the second<br />

station (Figure 22). This may be<br />

achieved by addressing either of<br />

the stations (destabilizing the initially<br />

preferred site or increasing the<br />

binding affinity of the originally<br />

weaker station). The system can be<br />

returned to its original state by using<br />

a second chemical modification to<br />

restore the initial order of station<br />

binding affinities. Performed consecutively,these<br />

two steps allow the<br />

“machine” to carry out a complete<br />

cycle of shuttling motion.<br />

The physical basis for this<br />

motion is again best understood by<br />

consideration of the potential<br />

energy of the macrocycle as a function<br />

of its position along the thread<br />

(Figure 23). It is important to appreciate that the external<br />

stimulus does not actually induce directional motion of the<br />

macrocycle,rather the system is put out of co-conformational<br />

equilibrium by increasing the binding strength of the lesspopulated<br />

station <strong>and</strong>/or destabilizing the initially preferred<br />

binding site. [215] Relaxation towards the new global energy<br />

minimum subsequently occurs by thermally activated motion<br />

of the components,a phenomenon which is recognized as<br />

biased Brownian motion. This effect can be envisaged as a net<br />

directional transport (a directional flux) of macrocycles<br />

towards the newly preferred station.<br />

106 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 22. Translational submolecular motion in a stimuli-responsive<br />

molecular shuttle: a) the macrocycle initially resides on the preferred<br />

station (orange); b) a reaction occurs (blue!green) which changes<br />

the relative binding potentials of the two stations such that, c) the<br />

macrocycle “shuttles” to the now-preferred station (green). If the<br />

reverse reaction (green!blue) now occurs (d), the components return<br />

to their original positions.<br />

Figure 23. Idealized potential energy of the macrocycle in a stimuli-responsive molecular shuttle in which<br />

one station changes (A’!A) in response to the stimulus <strong>and</strong> complicating factors such as folding are<br />

ignored. As before (Figure 21), the potential-energy surface shows the effect of the interaction between the<br />

macrocycle <strong>and</strong> thread on the energy of the macrocycle. The main plot shows the potential energy in terms<br />

of the position of the macrocycle along the vector of the thread; the minor plot shows the potential energy<br />

in terms of the position of the macrocycle orthogonal to the vector of the thread.<br />

Given this mode of action,a key requirement is finding<br />

ways of generating large long-lived binding energy differences<br />

between pairs of positional isomers. A Boltzmann distribution<br />

at 298 K requires a DDE (or DDG) value between<br />

translational co-conformers of about 2 kcalmol 1 for 95%<br />

occupancy of one station. Achieving such discrimination in<br />

two states to form a positionally bistable shuttle (that is,both<br />

DDE A’-B <strong>and</strong> DDE B-A 2 kcalmol 1 ) by modifying only<br />

intrinsically weak,noncovalent binding modes thus presents<br />

a significant challenge.<br />

4.3.3. Adding <strong>and</strong> Removing Protons to Induce Net Positional<br />

Change<br />

The first bistable switchable molecular shuttle,reported<br />

by Stoddart,Kaifer,<strong>and</strong> co-workers in 1994 <strong>and</strong> arguably the<br />

first true example of a molecular Brownian motion machine,<br />

Angew<strong>and</strong>te<br />

Chemie<br />

employed a two-station design (Scheme 34). [216] The biphenol<br />

<strong>and</strong> benzidine units in the thread of [2]rotaxane 68 4+ are both<br />

potential p-electron-donor stations for the cyclobis(paraquatp-phenylene)<br />

(CBPQT 4+ ) cyclophane,<strong>and</strong> at room temperature<br />

rapid shuttling of the macrocycle occurs as in related<br />

degenerate shuttles (Section 4.2.1). Cooling the solution to<br />

229 K allowed observation (by NMR <strong>and</strong> UV/Vis absorption<br />

spectroscopy) of the two translational isomers in a ratio of<br />

21:4 in favor of encapsulation of the benzidine station.<br />

Protonation of the benzidine residue with CF3CO2D destabilizes<br />

its interaction with the macrocycle so that it now<br />

resides overwhelmingly on the biphenol station. The system<br />

can be restored to its original state by neutralization with<br />

[D5]pyridine. [217]<br />

Although this system is a controllable molecular shuttle,it<br />

exhibits modest positional integrity in the nonprotonated<br />

state—a much larger binding<br />

difference between the two<br />

stations is required. Inspired<br />

by the complexation of ammonium<br />

ions by crown<br />

ethers,a new series of rotaxanes<br />

based on threads containing<br />

secondary alkylammonium<br />

species <strong>and</strong> crown<br />

ether macrocycles was developed.<br />

[218]<br />

[2]Rotaxane<br />

[69·H] 3+<br />

(Scheme 35) was<br />

the first switchable molecular<br />

shuttle reported using<br />

these units. [219] The dibenzocrown<br />

ether is bound to the<br />

ammonium cation by means<br />

of N + H···O hydrogen bonds<br />

as well as weaker C H···O<br />

bonds from methylene<br />

groups in the a position to<br />

the nitrogen atom. 1 H NMR<br />

spectroscopic analysis of the<br />

rotaxane in [D 6]acetone at<br />

room temperature shows (to<br />

the limits of detection of this<br />

experimental technique) the crown ether to be sitting mainly<br />

(but not exclusively [220] ) over the ammonium ion. Deprotonation<br />

of the ammonium center with diisopropylethylamine<br />

turns off the interactions holding the macrocycle to this<br />

station so it resides on the alternative bipyridinium station<br />

(69 2+ ). [221] A kinetic study of the shuttling in a closely related<br />

rotaxane revealed that the base-induced step is significantly<br />

slower than the return motion. This occurs despite a lower<br />

enthalpy of activation for the forward step <strong>and</strong> is due to<br />

entropic factors arising from the rearrangement of counterions<br />

in the transition state. [219c,222]<br />

While this example exhibits good macrocycle positional<br />

integrity in both chemical states,the low binding constant<br />

between the crown ether <strong>and</strong> bipyridinium moieties in the<br />

deprotonated state might be a limitation in more complex<br />

systems,especially if the macrocycle is afforded a greater<br />

degree of translational freedom (the distance between the<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

107


Reviews<br />

Scheme 34. The first switchable molecular shuttle 68 4+ /[68·2 H] 6+ . [216] The macrocycle<br />

distribution in 68 4+ was determined by 1 H NMR spectroscopy at 229 K in<br />

CD 3CN. The macrocycle distribution in [68·2H] 6+ is based on the lack of<br />

observation of the other translational isomer by 1 H NMR spectroscopy at a<br />

number of temperatures in the range 193–304 K in [D 6]acetone<br />

stations in the current system is ca. 7 Š). Accordingly,<br />

Stoddart,Balzani,<strong>and</strong> co-workers prepared a combination<br />

of three shuttle units in parallel by using this system. [223] The<br />

molecule ([70·3 H] 9+ ,Scheme 36) consists of three threadlike<br />

components connected at one end <strong>and</strong> encircled by three<br />

catechol-polyether macrocycles connected in a platformlike<br />

fashion. Essentially,the shuttling action works as in 69; just<br />

over three equivalents of base are required to move the<br />

platform from the ammonium (“top”) to the bipyridinium<br />

(“bottom”) positions. Titration experiments <strong>and</strong> molecularmodeling<br />

studies indicate that the shuttling motion proceeds<br />

in a stepwise fashion with each polyether macrocycle stepping<br />

individually from station to station. The effect of combining<br />

the three binding sites is excellent positional integrity in both<br />

the fully protonated <strong>and</strong> fully deprotonated states,while a<br />

further analogue utilizing a dioxynaphthalene trismacrocyclic<br />

platform exhibits even stronger interactions when on the<br />

bipyridyl station. [223b]<br />

The first pH-switchable shuttle to exploit hydrogenbonding<br />

interactions to anions was recently reported. [224] In<br />

[2]rotaxane 71·H (Scheme 37) formation of the benzylic<br />

amide macrocycle is templated by a succinamide station in<br />

the thread. However,the thread also contains a cinnamate<br />

derivative. In the neutral form the macrocycle resides on the<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

succinamide station more than 95% of the time<br />

because the phenol ring is a poor hydrogenbonding<br />

group. Deprotonation in [D 7]DMF to<br />

give 71 results in the macrocycle changing<br />

position to bind to the strongly hydrogen-bondbasic<br />

phenolate anion. Reprotonation returns the<br />

system to its original state <strong>and</strong> the macrocycle to<br />

its original position. However,the shuttling is<br />

extremely solvent-dependent. Hydrogen-bonding<br />

molecular shuttles usually perform best in “noncompeting”<br />

solvents. However,when the deprotonation<br />

of 71·H is carried out in CDCl 3 or<br />

CD 2Cl 2,a change of position of the macrocycle<br />

does not occur. Instead,intramolecular folding<br />

occurs to allow the phenolate group to hydrogen<br />

bond with the macrocycle while it remains on the<br />

succinamide station. This solvent dependence is<br />

explained by the fact that the phenolate group can<br />

only satisfy the hydrogen-bonding requirements<br />

of one isophthalamide unit of the macrocycle. The<br />

DMF facilitates shuttling by hydrogen bonding to<br />

the remaining amide groups of the macrocycle.<br />

Shuttling is unaffected by the nature of the<br />

accompanying cation or the addition of up to ten<br />

equivalents of other anions.<br />

Scheme 35. A pH-responsive bistable molecular shuttle displaying<br />

good positional integrity in both chemical states. [219] Statistical<br />

distributions of the macrocycle were determined at a number of<br />

temperatures in the range 193–304 K in [D 6]acetone.<br />

108 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 36. Chemical structure of a “molecular elevator” [70·3H] 9+ /70 6+ . [223]<br />

Scheme 37. pH-switched anion-induced shuttling in hydrogen-bonded<br />

[2]rotaxane 71·H/71 in [D 7]DMF at RT. [224] Bases that are effective<br />

include LiOH, NaOH, KOH, CsOH, Bu 4NOH, tBuOK, DBU, <strong>and</strong><br />

Schwesinger’s phosphazine P 1 base, thus illustrating the lack of<br />

influence of the accompanying cation on shuttling.<br />

4.3.4. Adding <strong>and</strong> Removing Electrons To Induce Net Positional<br />

Change<br />

In the original switchable shuttle 68 4+ (Scheme 34),the<br />

change of position could also be achieved in a reagent-free<br />

manner by electrochemical oxidation of the benzidine<br />

station; the macrocycle shuttles away from this station after<br />

oxidation to the radical cation. [216] A number of attempts to<br />

Angew<strong>and</strong>te<br />

Chemie<br />

create related redox-switched shuttles with greater positional<br />

integrity failed to give an improved distribution of translational<br />

isomers in the ground state. [225,226] Incorporation of<br />

redox-active stations based on tetrathiafulvalene (TTF),<br />

however,has given rise to a whole series of redox-active<br />

shuttles,a typical example of which (72 4+ /72 6+ ) is illustrated in<br />

Scheme 38. [227] Switching of the CBPQT 4+ cyclophane<br />

between TTF (or closely related derivatives) <strong>and</strong> dioxyarene<br />

units has since been extensively studied in both rotaxane <strong>and</strong><br />

catenane (see Section 4.6.1) architectures (as well as a<br />

number of model systems) to fully underst<strong>and</strong> the relationship<br />

between chemical structure <strong>and</strong> performance characteristics<br />

with a view to incorporate these shuttles in various<br />

condensed-phase devices (see Section 8.3). [180c,184l,u,186c,227,228]<br />

One of the key findings to emerge from these studies is that<br />

there is a significant activation barrier for the shuttling of the<br />

ring back onto the TTF unit when it is reduced from the<br />

oxidized state. This is in contrast to movement away from the<br />

freshly oxidized TTF unit which involves a very low kinetic<br />

barrier on account of the unfavorable interaction between the<br />

two positively charged components. The metastable state can<br />

be observed in solution at low temperature, [228f] <strong>and</strong> its<br />

characterization has been crucial in determining the mechanism<br />

of operation of the solid-state electronic devices<br />

constructed from molecules of this type (see Section 8.3.1).<br />

An electrochemical switch with one of the highest<br />

positional fidelities (ca. 10 6 :1 in one state; ca. 1:500 in the<br />

other) has been demonstrated with amide-based molecular<br />

shuttles. [229] [2]Rotaxane 73 contains two potential hydrogenbonding<br />

stations (a succinamide (succ) station <strong>and</strong> a redoxactive<br />

3,6-di-tert-butyl-1,8-naphthalimide (ni) station) for the<br />

benzylic amide macrocycle that are separated by a C 12<br />

aliphatic spacer (Scheme 39). [229]<br />

While the ability of the succ station to template formation<br />

of the macrocycle is well established,the neutral naphthalimide<br />

moiety is a poor hydrogen-bond acceptor. In fact,the<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

109


Reviews<br />

Scheme 38. Redox-switchable molecular shuttle 72 4+ /72 6+ . [227] The redox reactions may be carried out either chemically or electrochemically.<br />

Scheme 39. A photochemically <strong>and</strong> electrochemically switchable, hydrogen-bonded<br />

molecular shuttle 73. [229] In the neutral state, the translational co-conformation succ-<br />

73 is predominant as the ni station is a poor hydrogen-bond acceptor<br />

(K n = (1.2 1) ” 10 6 ). Upon reduction, the equilibrium between succ-73 C <strong>and</strong><br />

ni-73 C is altered (K red = (5 1) ” 10 2 ) because ni C is a powerful hydrogen-bond<br />

acceptor <strong>and</strong> the macrocycle moves through biased Brownian motion. Upon<br />

reoxidation, the macrocycle shuttles back to the succinamide station. Repeated<br />

reduction <strong>and</strong> oxidation causes the macrocycle to shuttle forwards <strong>and</strong> backwards<br />

between the two stations. All the values shown refer to cyclic voltammetry experiments<br />

in anhydrous THF at 298 K with tetrabutylammonium hexafluorophosphate<br />

as the supporting electrolyte. Similar values were determined on photoexcitation<br />

<strong>and</strong> reduction of the ensuing triplet excited state by an external electron donor.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

difference in macrocycle binding affinities is so great that<br />

succ-73 is the only translational isomer detectable by<br />

1 H NMR specroscopy in CDCl3,CD 3CN,<strong>and</strong> [D 8]THF (an<br />

equilibrium constant of (1.2 1) ” 10 6 for the two translational<br />

isomers was determined by electrochemistry in<br />

THF containing Bu 4NPF 6 as a supporting electrolyte);<br />

even in the strongly hydrogen-bond-disrupting<br />

[D 6]DMSO,the macrocycle resides over the succ station<br />

about half of the time. One-electron reduction of naphthalimide<br />

to the corresponding radical anion,however,<br />

results in a substantial increase in electron charge density<br />

on the imide carbonyl groups <strong>and</strong> a concomitant increase<br />

in hydrogen-bond-accepting ability. In 73,this reverses the<br />

relative hydrogen-bonding abilities of the two thread<br />

stations so that co-conformation ni-73 C is preferred over<br />

succ-73 C to the extent of approximately 500:1. Subsequent<br />

reoxidation to the neutral state restores the original<br />

binding affinities <strong>and</strong> the shuttle returns to its initial<br />

state. The process can be observed in cyclic voltammetry<br />

experiments, [229b] or alternatively photochemistry can be<br />

employed to initiate (through excitation of the naphthalimide<br />

group by a nanosecond laser pulse at 355 nm<br />

followed by electron transfer from an external electron<br />

donor) <strong>and</strong> observe (by using transient absorption spectroscopy)<br />

the change of position. [229a] Importantly,a<br />

number of control experiments involving the incorporation<br />

of steric barriers to shuttling were carried out to prove<br />

unequivocally that the dynamic process observed in this<br />

rotaxane system is a reversible shuttling of the macrocycle<br />

between the stations rather than any other conformational<br />

or co-conformational changes. [229b,230]<br />

Although many metal–lig<strong>and</strong> interactions can be<br />

rather kinetically stable,the difference in preferred<br />

coordination geometries of different oxidation states can<br />

be exploited to bring about redox-induced shuttling in<br />

transition-metal-based systems. The preference of Cu I ions<br />

for four-coordinate tetrahedral complexes has been widely<br />

employed in the synthesis of interlocked architectures, [189-<br />

110 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

g,195a,e–g,j,af]<br />

as it enforces the orthogonal arrangement of two<br />

bidentate lig<strong>and</strong>s which can be further elaborated to give<br />

rotaxanes or catenanes. In [Cu I (74)] (Scheme 40),a macrocycle<br />

containing a bidentate 2,9-diphenyl-1,10-phenanthroline<br />

(dpp) unit,is locked onto a thread which contains one<br />

bidentate phenanthroline (phen) <strong>and</strong> one tridentate 2,2’:6’,2’’terpyridine<br />

(terpy) unit. [231] Chelation of Cu I ions ensures that<br />

the macrocycle is positioned over the phenanthroline station<br />

in the thread,phen-[Cu I (74)]. Electrochemical oxidation<br />

initially generates phen-[Cu II (74)]. However,as Cu II ions<br />

prefer higher coordination numbers,a thermally activated<br />

relaxation to the thermodynamically preferred terpy-<br />

[Cu II (74)] occurs. The metastable phen-[Cu II (74)] state is<br />

relatively kinetically inert,however,<strong>and</strong> this shuttling step<br />

takes several hours at room temperature (k = 1.5 ” 10 4 s 1 ).<br />

Electrochemical reduction of the divalent species gives terpy-<br />

[Cu I (74)] which slowly converts into the original translational<br />

isomer phen-[Cu I (74)] (10 4<br />

k 10 2 s 1 ,which is slightly<br />

faster than the first step because of the smaller charge on the<br />

metal center). The same shuttling process can also be<br />

accomplished by a photochemically triggered oxidation. The<br />

reverse reduction step cannot be carried out photochemically<br />

but proceeds successfully with a chemical reductant (ascorbic<br />

acid). [232]<br />

Scheme 40. Redox-switched shuttling in the metal-templated [2]rotaxane 74. [231]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

4.3.5. Adding <strong>and</strong> Removing Metal Ions To Induce Net Positional<br />

Change<br />

The research groups of S<strong>and</strong>ers <strong>and</strong> Stoddart have<br />

collaborated to produce another class of molecular shuttles<br />

that can be switched in several ways (Scheme 41). In the<br />

ground state,the co-conformation of kinetically stable<br />

[2]pseudorotaxane 75 with the macrocycle sitting over the<br />

naphthaldiimide station (blue) is dominant,as observed by<br />

1 H NMR spectroscopy <strong>and</strong> one-electron reduction of this<br />

station (which has the lower reduction potential) results in<br />

shuttling of the ring onto the pyromellitic diimide station<br />

(green). Somewhat unexpectedly,however,the change in<br />

position of the macrocycle can also be induced by adding<br />

lithium ions. Two of the small metal ions can be complexed<br />

between the crown ether macrocycle <strong>and</strong> the carbonyl groups<br />

of the diimide moieties <strong>and</strong> this interaction is significantly<br />

stronger at the pyromellitic station (Scheme 41). Addition of<br />

a large excess of [18]crown-6 sequesters the lithium ions,<br />

thereby returning the system to its initial state. [233]<br />

Rather than serving to enhance intercomponent interactions<br />

at a specific station,metal-ion binding can be used to<br />

destabilize the binding of the macrocycle at one site,thus<br />

causing it to translocate to another,unaffected unit. This is<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

111


Reviews<br />

Scheme 41. Cation-induced shuttling based on complexation <strong>and</strong> decomplexation of lithium ions. [233]<br />

possible through two distinct mechanisms (Scheme 42). [234,235]<br />

In [2]rotaxane 76 the macrocycle preferentially sits over the<br />

glycylglycine station derivatized with a bis(2-picolyl)amino<br />

(BPA) stopper. The addition of one equivalent of Cd-<br />

(NO3) 2·4 H2O generates a complex in which the metal ion is<br />

bound to the first carboxamide carbonyl oxygen atom as well<br />

as the three nitrogen donors of the BPA lig<strong>and</strong>,<strong>and</strong> the<br />

preferred postion of the macrocycle remains essentially<br />

unchanged. However,deprotonation of the first carboxamide<br />

moiety [236] results in coordination of the nitrogen anion,as<br />

well as the carbonyl oxygen atom of the second carboxamide<br />

group. The cadmium ion essentially wraps itself up in the<br />

deprotonated glycylglycine residue,thus switching off any<br />

intercomponent interactions with the macrocycle which now<br />

occupies the succinic amide ester station. The shuttling<br />

process is fully reversible: removal of the Cd II<br />

ion with<br />

excess cyanide <strong>and</strong> reprotonation of the amide nitrogen atom<br />

with NH4Cl quantitatively regenerates 76. [234] In [2]rotaxane<br />

77,again the macrocycle preferentially resides on the<br />

carboxamide-based station adjacent to a BPA lig<strong>and</strong>. In this<br />

case,divalent metal ions such as Cd II are only able to chelate<br />

to the three BPA nitrogen atoms. To accommodate such a<br />

binding mode,however,the two pyridyl arms must adopt a<br />

coplanar conformation,which sterically destabilizes macrocycle<br />

binding to the succinamide station,thus causing it to<br />

move to the inherently weaker succinic amide ester unit.<br />

Rather than competing for the same donor atoms,as in 76,the<br />

metal- <strong>and</strong> macrocycle-binding modes in 77 compete for the<br />

same 3D space so that this mechanism corresponds to a<br />

negative heterotropic allosteric binding event. The shuttling is<br />

fully reversible on demetalation of [Cd(NO3) 2(77)] with<br />

cyanide. [235]<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

4.3.6. Adding <strong>and</strong> Removing Covalent Bonds To Induce Net<br />

Positional Change<br />

Perhaps surprisingly,the use of covalent-bond-forming<br />

reactions to bring about positional change in molecular<br />

shuttles has not yet been explored extensively. In one<br />

successful example,the formation (<strong>and</strong> breaking) of C C<br />

bonds through Diels–Alder <strong>and</strong> retro-Diels–Alder reactions<br />

of rotaxane 78 (Scheme 43) can control shuttling with<br />

excellent positional discrimination: the steric bulk of the<br />

Diels–Alder adduct displaces the macrocycle to the succinic<br />

amide ester station in Cp-78. [237]<br />

A transient shuttling system involving photoinduced<br />

cleavage of a covalent bond <strong>and</strong> subsequent recombination<br />

has recently been reported. Diaryl cycloheptatrienes can act<br />

as p-electron-donor binding sites in rotaxanes containing<br />

CBPQT 4+ macrocycles. [180b,d] In [2]rotaxane 79 4+ ,the cationic<br />

cyclophane resides over the diarylcycloheptatriene station<br />

(green) with additional “alongside” interactions with the<br />

anisole unit (red). Photoheterolysis of the C OMe bond<br />

under conventional flash photolysis conditions at 360 nm<br />

generates the diaryl tropylium cation <strong>and</strong> displaces the ring to<br />

the anisole station (Scheme 44). The lifetime of this species at<br />

room temperature is 15 s,before the thermal back reaction<br />

returns the system to 79 4+ (or a regioisomer resulting from<br />

nucleophilic attack of the methoxide at a different site on the<br />

seven-membered ring). The shuttling motion was inferred by<br />

comparison of the transient absorption spectrum of<br />

79 5+ ·MeO with the spectral characteristics of 79 5+ generated<br />

by electrochemical oxidation. [238] Recently,modifications to<br />

the inactive thread components have allowed preparation of<br />

an analogue which adopts a linear conformation in the ground<br />

112 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 42. a) Shuttling through stepwise competitive binding. [234] A<br />

similar shuttling mechanism can be observed with Cu II ions, where the<br />

deprotonation results in a color change. b) An allosterically regulated<br />

molecular shuttle. [235]<br />

Scheme 43. Shuttling through reversible formation of covalent<br />

bonds. [237] The absolute stereochemistry for Cp-78 is depicted<br />

arbitrarily.<br />

Scheme 44. Photoheterolysis-induced reversible shuttling through<br />

formation of a tropylium cation. [238]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

state <strong>and</strong> for which formation of the tropylium cation can be<br />

achieved with acids,as well as through the photoheterolysis<br />

mechanism. [239]<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

113


Reviews<br />

4.3.7. Changing Configuration To Induce Net Positional Change<br />

As with controlling motion in covalently bonded systems,<br />

isomerization processes,particularly photoisomerization processes,are<br />

a very attractive means of inducing shuttling in<br />

rotaxanes. Shuttle (E)/(Z)-80 (Scheme 45) utilizes the inter-<br />

Scheme 45. Bistable molecular shuttle (E)/(Z)-80 in which self-binding<br />

of the low affinity station in each state is a major factor in producing<br />

excellent positional discrimination. [240] Similar results are achieved<br />

when the intermediate affinity station (orange) is a succinic amide<br />

ester.<br />

conversion of fumaramide (trans) <strong>and</strong> maleamide (cis)<br />

isomers of the olefinic unit. [240] Fumaramide groups are<br />

excellent binding sites for benzylic amide macrocycles: [241]<br />

the trans olefin holds the two strongly hydrogen bond accepting<br />

amide carbonyl groups in a preorganized close-to-ideal<br />

spatial arrangement for interaction with the amide groups of<br />

the macrocycle. Although a similar hydrogen-bonding surface<br />

is presented to the host,binding of the macrocycle to the<br />

succinamide station in (E)-80 would result in a loss in entropy<br />

(loss of bond rotation in the succinamide group) as well as one<br />

less intracomponent hydrogen bond than would be present in<br />

the fumaramide-occupied positional isomer. The result is that<br />

only one major positional isomer of (E)-80 (shown in<br />

Scheme 45) is observed at room temperature in CDCl 3.<br />

Photoisomerization (254 nm) reduces the number of possible<br />

intercomponent hydrogen bonds at the olefin station from<br />

four to two <strong>and</strong> so the macrocycle changes position to the<br />

succinamide station ((Z)-80,Scheme 45). Unlike many other<br />

light-switchable shuttles,this new state is essentially indefinitely<br />

stable until a further stimulus is applied to reisomerize<br />

the maleamide unit back to fumaramide.<br />

4.3.8. Shuttling through Excited States<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

When operating in the light-stimulated mode,[2]rotaxane<br />

73 (Section 4.3.4) employs a highly oxidizing photochemically<br />

excited state to trigger the electron-transfer process which<br />

results in macrocycle shuttling. After about 100 ms,the<br />

reduced rotaxane undergoes charge recombination with the<br />

radical cation of the external electron donor to regenerate the<br />

starting state. As long as photons of a suitable wavelength are<br />

supplied,this process will continue to occur indefinitely; the<br />

photochemical process can be said to be autonomous. The<br />

same principle is also true of shuttle 79 (Section 4.3.6).<br />

Attempts to induce shuttling using intramolecular electron<br />

transfer from photoexcited states have in the past been<br />

plagued by the problem of back electron transfer being fast<br />

compared to the desired nuclear motion. [207,232b] Recently,<br />

however,it has been demonstrated that a previously studied<br />

[2]rotaxane, 81 6+ , [232b] does indeed undergo shuttling in an<br />

intramolecular charge-separated state (Figure 24). [242,3] Irradiation<br />

of the ruthenium–trisbipyridine complex (green)<br />

generates a highly reducing excited state. An intramolecular<br />

electron transfer then occurs between the excited metal<br />

center <strong>and</strong> the most easily reduced bipyridinium station<br />

(blue),on which the macrocycle prefers to sit. The result is<br />

destabilization of the macrocycle–station interactions so that<br />

the alternative bipyridinium unit (pink) is now preferred.<br />

Remarkably,in this system the back electron-transfer process<br />

is slow enough to allow shuttling of the ring towards the other<br />

station in approximately 10 % of the molecules. Naturally,<br />

charge recombination quickly restores the initial state,but if<br />

photons are continually supplied,this cycle is followed<br />

indefinitely.<br />

In another example,shuttling has been observed simply as<br />

a result of the altered electron density in a photoexcited state,<br />

without any subsequent electron-transfer reactions. The nonplanarity<br />

of the anthracene-9-carboxamide unit in [2]rotaxane<br />

82 means the macrocycle can only form hydrogen bonds<br />

with one of the amide carbonyl groups (Scheme 46). [243]<br />

Photoexcitation of the anthracene system causes a planar<br />

conformation to be adopted in which some charge is transferred<br />

onto the adjacent carbonyl group,thus increasing its<br />

hydrogen-bond basicity. Subsequent translocation of the<br />

macrocycle occurs on a nanosecond timescale <strong>and</strong> is reflected<br />

in a change in the anthracene fluorescence spectrum,before<br />

decay of the excited state restores the original co-conformation.<br />

These shuttles operate without the consumption of<br />

chemical fuels or the formation of waste products <strong>and</strong> they<br />

automatically reset so that they operate autonomously. It<br />

must be noted,however,that if a continuous supply of<br />

photons is provided the distribution of the rings between the<br />

two stations would reach a steady state (the exact ratio<br />

depending on the intensity of the light) within a few milliseconds.<br />

For an autonomous molecular motor (for example,<br />

46; see Section 2.2) a steady state of isomer populations can<br />

still correspond to a net flux in a particular direction through<br />

these forms. In switches such as these shuttles,however,after<br />

the steady state has been reached,there is no subsequent net<br />

flux of rings between the two stations at any point in time. The<br />

114 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 24. Chemical structure (a) <strong>and</strong> operating cycle in a schematic form (b) of a molecular shuttle 81 6+<br />

operating through a photoinduced internal charge-separated state. [242,3] At equilibrium in the ground state,<br />

the ring spends most of the time over the unsubstituted bipyridinium station (blue, A). Irradiation (1) of the<br />

ruthenium complex (green) generates a highly reducing excited state, which results in electron transfer (2)<br />

to the blue station, thus weakening its electrostatic interactions with the ring (B). Normally charge<br />

recombination processes such as (3) are fast in comparison with nuclear motions, but here it is slow<br />

enough to allow approximately 10 % of the molecules to undergo significant Brownian motion (4), thereby<br />

shifting the statistical distribution in this portion of the ensemble to favor the dimethylbipyridinium station<br />

(pink, C). When charge recombination (5) eventually does take place, the higher binding affinity of the blue<br />

station is restored (D). The system relaxes (6) to restore the original statistical distribution of rings (A).<br />

Scheme 46. Photoinduced shuttling in the singlet excited state of<br />

benzylic amide [2]rotaxane 82. [243] The shuttling amplitude is approximately<br />

3 Š <strong>and</strong> the lifetime of 82* is 4.3 ns.<br />

only way to generate net<br />

fluxes of macrocycles<br />

between the stations in<br />

these types of systems<br />

would be to rapidly switch<br />

the photon source on <strong>and</strong><br />

off. [3]<br />

4.3.9. Entropy-Driven<br />

Shuttling<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Most of the shuttles<br />

which exhibit excellent positional<br />

discrimination<br />

between two stations are<br />

switched using stimuli to<br />

modify the enthalpy of the<br />

macrocycle binding to one or<br />

both stations. Generally,the<br />

effect of temperature is only<br />

to alter the degree of discrimination,not<br />

to alter the<br />

station preference. [244]<br />

In<br />

[2]rotaxane 83,however,the<br />

macrocycle can be switched<br />

between stations simply by<br />

changing the temperature.<br />

[245] In fact, 83 is a tristable<br />

molecular shuttle: the<br />

ring can be switched<br />

between three different positions<br />

on the thread<br />

(Scheme 47).<br />

Structurally, 83 is closely<br />

related to 80,the differences<br />

being substitution of the isophthaloyl<br />

unit in the macrocycle for pyridine-2,6-dicarbonyl<br />

groups <strong>and</strong> replacement of the succinamide station with a<br />

succinic amide ester. In the (E)-83 form,the macrocycle<br />

resides over the fumaramide station in CDCl 3 at all temperatures<br />

investigated. However,although the 1 H NMR spectrum<br />

of the maleamide isomer ((Z)-83) in CDCl 3 showed that<br />

the macrocycle was no longer positioned over the olefin<br />

station,the spectrum was highly temperature-dependent. At<br />

elevated temperatures (308 K) the expected succ-(Z)-83 coconformation<br />

was observed,but at lower temperatures it is<br />

the alkyl chain which exhibits the spectroscopic shifts<br />

indicative of encapsulation by the macrocycle (dodec-(Z)-<br />

83,Scheme 47). The origin of this temperature-switchable<br />

effect is the large difference in the entropy of binding<br />

(DS binding) to the succinamide <strong>and</strong> alkyl chain stations which<br />

allows the TDS binding term to have a significant impact on the<br />

DG binding value as the temperature is varied. [246] In the succ-<br />

(Z)-83 co-conformation,the macrocycle forms two strong<br />

hydrogen bonds with an amide carbonyl group <strong>and</strong> two,<br />

significantly weaker,bonds to the ester carbonyl group. The<br />

dodec-(Z)-83 co-conformation allows the formation of four<br />

strong hydrogen bonds to amide carbonyl groups,thus<br />

making it enthalpically favored by about 2 kcal mol 1 . At<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

115


Reviews<br />

Scheme 47. A tristable molecular shuttle 83. [245]<br />

low temperatures,where the effects of entropy are less<br />

significant,the molecule adopts the dodec-(Z)-83 co-conformation.<br />

At higher temperatures,the increased contribution<br />

from the TDS binding term favors the entropically preferred<br />

succ-(Z)-83 co-conformation. [247]<br />

4.3.10. Shuttling through a Change in the Nature of the<br />

Environment<br />

A significant advantage [248] of using temperature to induce<br />

a change in net position of the macrocycle in a molecular<br />

shuttle is that no chemical reaction is involved <strong>and</strong> any<br />

property change associated with the new state cannot arise<br />

from a change in the covalent structure of the molecule. The<br />

same holds true for switching induced by a change in the<br />

nature (solvation or polarity) of the environment of the<br />

shuttle. The benzylic amide macrocycle in amphiphilic<br />

rotaxanes (including those discussed in Section 4.2.1) can be<br />

shuttled between various hydrogen-bonding stations (CDCl 3<br />

<strong>and</strong> most nonpolar solvents) <strong>and</strong> an alkyl chain ([D 6]DMSO,<br />

H 2NCHO). [199,249] A similar effect has been observed in a<br />

range of poly(urethane/crown ether rotaxane)s. [250]<br />

Many of the examples of stimuli-induced shuttling described<br />

in Sections 4.3.1–4.3.9 display remarkable degrees of<br />

control over the positioning <strong>and</strong> dynamics of submolecular<br />

fragments. They utilize a number of different stimuli-induced<br />

processes to trigger changes in the net position of macrocycles<br />

over large distances (up to ca. 15 Š in the case of 71, 73,<strong>and</strong><br />

80) <strong>and</strong> operate over a range of timescales (a complete<br />

switching cycle in 73 is over in ca. 100 ms while,in 80,both<br />

states are effectively indefinitely stable). Note,however,that<br />

all these shuttles exist as an equilibrium of co-conformations<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

<strong>and</strong> it is just the position of the<br />

equilibrium that is being varied. [220]<br />

As the position of equilibrium<br />

changes,detailed balance is broken<br />

<strong>and</strong> it is this that can allow a useful<br />

mechanical task to be performed by<br />

Brownian motion of the macrocycle.<br />

4.4. Compartmentalized <strong>Molecular</strong><br />

<strong>Machines</strong><br />

We discussed in Section 4.3.2 how<br />

stimuli-responsive molecular shuttles<br />

can be considered in terms of a<br />

machine (the thread) which transports<br />

a particle (the interlocked macrocycle)<br />

between two sites (compartments)<br />

on a one-dimensional potential-energy<br />

surface. This approach<br />

provides a basic framework with<br />

which to explore some of the ideas<br />

to create,control,order,<strong>and</strong> manipulate<br />

non-equilibrium conformations<br />

<strong>and</strong> co-conformations raised at the<br />

end of Section 1.4.2. In all of the<br />

examples in Section 4.3,an external<br />

stimulus alters the relative binding affinities of the stations for<br />

the macrocycle by placing the system momentarily out of coconformational<br />

equilibrium. Detailed balance is broken <strong>and</strong><br />

the system acts to restore balance through biased Brownian<br />

motion of the macrocycles towards the new global minimum.<br />

Detailed balance,however,can be considered to result from<br />

two separate properties of the system (Section 1.4.1): the<br />

statistical distribution of a quantity (an imbalance which<br />

provides the thermodynamic impetus for net transport) <strong>and</strong><br />

the ability of that quantity to be dynamically exchanged<br />

(which provides the communication necessary for transport to<br />

occur). [51] In the switchable shuttles of Section 4.3,the macrocycle<br />

is at all times able to seek its equilibrium statistical<br />

distribution along the full length of the thread,<strong>and</strong> the<br />

position of the substrate is always indicated by the state of the<br />

machine. To create artificial Brownian machines which are<br />

more sophisticated than simple positional switches,control<br />

over the kinetics for exchange of the substrate between two<br />

sites of the machine must be introduced.<br />

Rotaxane system 84 is able to change the average position<br />

of the macrocycle irreversibly through a fundamentally<br />

different mechanism to previous shuttles (Scheme 48). [32] In<br />

84 the two stations are structurally identical (but distinguishable:<br />

note the different stoppers) <strong>and</strong> a bulky silyl ether acts<br />

as a barrier which prevents the ring from moving between<br />

them. The system starts out statistically unbalanced (succ1-84,<br />

because of the synthetic route used to access it). The stimulus<br />

that leads to the net change in position of the macrocycle is a<br />

“linking” operation—removal of the silyl ether—which<br />

switches “on” dynamic exchange of the macrocycle between<br />

the two stations <strong>and</strong> allows the system to move towards<br />

equilibrium,which results in an average displacement of the<br />

116 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 48. Operation of a compartmentalized Brownian<br />

molecular machine that acts as an irreversible switch. [32] As the<br />

macrocycle distributions on the thread are determined when<br />

the two compartments are unlinked, they are independent of<br />

factors such as temperature or solvent. TBDMS = tert-butyldimethylsilyl.<br />

macrocycle half the distance separating the two<br />

stations. This is a molecular-level form of “escapement”,the<br />

element of the mechanism that controls the<br />

release of potential energy to drive mechanical motion<br />

in clocks <strong>and</strong> other macroscopic mechanical devices.<br />

Raising the barrier by applying an “unlinking stimulus”<br />

resets the machine (the thread) without undoing<br />

the task just performed. However,applying the linking<br />

stimulus again after the machine is reset,does not<br />

change the average position of the macrocycle a<br />

second time,because the system is now statistically<br />

balanced. This stimuli-induced irreversible net change<br />

in the position of the macrocycle represents a new type<br />

of molecular shuttle in phenomenological terms: its<br />

operation is irreversible <strong>and</strong> the state of the machine<br />

(the thread) does not determine the position of the<br />

substrate.<br />

Combining control over exchange between the two<br />

stations (kinetics) with the ability to modulate their<br />

relative binding affinities (thermodynamics) led to the<br />

creation of another novel type of simple molecular<br />

machine system,namely 85 (Scheme 49). [32] Here,the<br />

machine is statistically balanced (85 % of the macrocycles<br />

on the fumaramide (green) station,15 % on the<br />

succinamide (orange) station),<strong>and</strong> unlinked (<strong>and</strong><br />

therefore not in equilibrium) by simply removing <strong>and</strong><br />

reattaching the silyl group (Scheme 49,steps a <strong>and</strong> c).<br />

From this balanced state,a balance-breaking stimulus<br />

is applied (irradiation at 312 nm,generating a 49:51<br />

2% E:Z photostationary state),followed by removal<br />

of the kinetic barrier (“linking stimulus”) which allows<br />

balance to be restored by biased Brownian motion of<br />

Angew<strong>and</strong>te<br />

Chemie<br />

the ring towards the new equilibrium distribution. Restoring<br />

the barrier (“unlinking stimulus”) makes the system unlinked<br />

<strong>and</strong> not in equilibrium,although statistically balanced. The<br />

resetting step (a different balance-breaking stimulus,to<br />

promote the Z!E olefin isomerization),makes the system<br />

statistically unbalanced,unlinked,<strong>and</strong> not in equilibrium.<br />

After the operational cycle of the machine,approximately<br />

56% of the macrocycles are located on the succinamide<br />

station. The transportation of the macrocycle in 85 is<br />

repeatedly reversible between the statistically balanced<br />

(85:15) <strong>and</strong> statistically unbalanced (44:56) ratios of fum-<br />

(E)-85 to succ-(E)-85.<br />

In this rotaxane the thread performs the task of directionally<br />

changing the net position of the macrocycle—<strong>and</strong><br />

since the succinamide station binds the macrocycle more<br />

weakly than the fumaramide station,the thread moves the<br />

macrocycle energetically uphill—while itself returning to its<br />

Scheme 49. Operation of compartmentalized molecular machine 85 which corresponds<br />

to a two-state Brownian flip-flop. [32] Operation steps: a) Desilylation (80% aqueous acetic<br />

acid); b) E!Z photoisomerization (hn at 312 nm); c) resilylation (TBDMSCl); <strong>and</strong><br />

d) Z!E thermal isomerization (catalytic piperidine). As the macrocycle distributions on<br />

the thread are determined when the two compartments are unlinked, they are<br />

independent of factors such as temperature or solvent.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

117


Reviews<br />

initial state. This behavior amounts to “ratcheting”,a<br />

characteristic feature of the operating mechanisms of many<br />

biological molecular machines. A thermodynamically unfavorable<br />

concentration gradient of the macrocycle between the<br />

compartments is the result,which is precisely the function<br />

envisaged for the thought-machine pressure demons discussed<br />

in Section 1.2.1. Unlike the thought machines,however,the<br />

position of the Brownian particle in 85 has no role in<br />

the mechanism. [251] Rather the rotaxane machine carries out a<br />

sequence of four steps (independent of the position of the<br />

particle) which govern in turn the thermodynamics <strong>and</strong> the<br />

kinetics for transport between the two stations (Figure 25):<br />

balance-breaking 1,linking,unlinking,balance-breaking 2<br />

(resetting,of the machine not the substrate).<br />

Figure 25. The operation [32] of machine–substrate system 85 in Scheme 49 is the experimental<br />

realization (albeit in non-adiabatic form) of the transportation task required of Smoluchowski’s<br />

trapdoor [16] (Figure 4a) <strong>and</strong> Maxwell’s pressure demon [15c] (Figure 2b). The mechanistic<br />

equivalent of the schematic representations from Section 1.4.2 is shown above. The colors of<br />

the compartments, particles, <strong>and</strong> door are the same as the corresponding elements of 85.<br />

The initially balanced (in proportion to the sizes of the two compartments) distribution of the<br />

Brownian particles between the left (L) <strong>and</strong> right (R) compartments becomes statistically<br />

unbalanced by a change in the volume of the left-h<strong>and</strong> compartment. Opening the door<br />

allows the particles to redistribute themselves according to the new size ratio of the<br />

compartments. Closing the door ratchets the new distribution of particles. Restoring the lefth<strong>and</strong><br />

compartment to its original size then results in a concentration gradient of the<br />

Brownian particles across the two compartments. There is no role for an informationgathering<br />

demon in this mechanism.<br />

Significantly,examining the state of the machine (the<br />

rotaxane thread) in 85 does not provide information regarding<br />

the state (distribution) of the substrate. It is only from the<br />

history of the machine s operations that the distribution of the<br />

substrate can be known; in other words,the net position of the<br />

macrocycle in rotaxane 85 is a consequence of a form of<br />

sequential logic. [252] The behavior of 85 is characteristic of a<br />

two-state (one-bit) memory or “flip-flop” component in<br />

electronics [253] <strong>and</strong> therefore 85 is the first example of a new<br />

class of simple molecular machine: a two-state Brownian flipflop.<br />

A flip-flop maintains its effect on a system indefinitely<br />

until an input pulse operates on it that causes its output to<br />

change to a new indefinitely stable state according to defined<br />

rules. [254]<br />

From the analysis of the behavior of these rotaxanes,<strong>and</strong><br />

the analysis of the workings of the fluctuation-driven transport<br />

mechanisms discussed in Section 1.4.2,there appear to<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

be four phenomenological terms (ratcheting,escapement,<br />

balance,<strong>and</strong> linkage) that are crucial for controlling the nonequilibrium<br />

distribution of Brownian substrates with compartmentalized<br />

molecular-level machines:<br />

1) “Ratcheting” [32]<br />

is an often used,but previously illdefined,process<br />

in chemical terms. Unfortunately,this<br />

vagueness has led to the term sometimes being applied to<br />

describe phenomena that are unrelated to Brownian<br />

ratchet mechanisms. Ratcheting is the capturing of a<br />

positional displacement of a substrate through the imposition<br />

of a kinetic energy barrier which prevents the<br />

displacement being reversed when the thermodynamic<br />

driving force is removed. The key feature of ratcheting is<br />

that the ratcheted part of the system is not linked with<br />

(that is,not allowed to exchange the substrate<br />

with) any part of the system that it is ratcheted<br />

from. Ratcheting is a crucial requirement for<br />

allowing a Brownian machine to be reset<br />

without undoing the task it has performed;<br />

2) “Escapement” [32] is the counterpart to ratcheting<br />

<strong>and</strong> consists of the (directional) release of a<br />

ratcheted substrate in a statistically unbalanced<br />

system by lowering a kinetic energy barrier (by<br />

linking). The key feature of escapement is that<br />

it requires the linking of two unbalanced parts<br />

of a system that were previously unlinked. An<br />

escapement step must be subsequently ratcheted<br />

for a machine to be able to do work<br />

repetitively on a substrate;<br />

3) “Balance” [32,51] is the thermodynamically preferred<br />

distribution of a substrate over a<br />

machine or parts of a machine. The impetus<br />

for net transportation of a substrate between<br />

two parts of a machine comes from the balance<br />

being broken (note that balance being broken<br />

is not the same as detailed balance being<br />

broken,balance is the thermodynamic driving<br />

force for detailed balance);<br />

4) “Linkage” [32,51] is the communication necessary<br />

for transportation of a substrate to occur<br />

between parts of a machine. However,the<br />

ability to exchange the substrate between the<br />

linked parts is not in itself enough for a task to be<br />

performed,there must also be a driving force for it to<br />

occur (see above). Linking <strong>and</strong> unlinking operations are<br />

purely kinetic parameters <strong>and</strong> so can be accomplished by<br />

simply changing the rate of reactions rather than introducing<br />

or removing physical barriers.<br />

The statistical balance of a dynamic substrate <strong>and</strong> whether<br />

the parts of the machine acting on the substrate allow<br />

exchange of the substrate or not appear to be key factors that<br />

determine whether the machine can perform a useful task or<br />

not. In fact,it appears that the behavior of a molecular<br />

machine towards a substrate can be defined by the changing<br />

relationship (linked/unlinked,balanced/unbalanced) between<br />

the parts of machine interacting with the substrate.<br />

We can see that there are three fundamental types of<br />

Brownian machine that act through various combinations of<br />

118 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

balance-breaking,linking/unlinking,ratcheting,<strong>and</strong> escapement<br />

steps: Brownian switches (for example,the switchable<br />

molecular shuttles in Section 4.3),Brownian flip-flops (for<br />

example,rotaxane 85),<strong>and</strong> Brownian motors (for example,17<br />

or 46,see Sections 2.1.2 <strong>and</strong> 2.2). By combining these simple<br />

Brownian machine types with other combinational <strong>and</strong>/or<br />

sequential operations based on Boolean logic,machines of<br />

increasing complexity can be envisaged (Figure 28).<br />

1) A “Two-state (or multistate) Brownian switch” [32] is a<br />

machine that can reversibly change the distribution or<br />

position of a Brownian substrate between two (or more)<br />

distinguishable sites as a function of the state of the<br />

machine. It does this by biasing the Brownian motion of<br />

the substrate (Figure 26 a). Classic stimuli-responsive<br />

Figure 26. Schematic representations of some simple compartmentalized<br />

molecular-level machines. a) Two-state Brownian switch. b) Irreversible<br />

Brownian switch. c) A two-state Brownian flip-flop (shown<br />

operating through a partial energy ratchet mechanism; other mechanisms<br />

can achieve the same machine function). [32]<br />

molecular shuttles,such as those discussed in Section 4.3,<br />

are examples of two-state (or three-state,Section 4.3.9)<br />

Brownian switches. An “irreversible Brownian switch” is a<br />

“once only” machine,such as succ1-84 (Scheme 48),which<br />

irreversibly changes the distribution or position of a<br />

Brownian substrate in response to an external stimulus<br />

(Figure 26 b).<br />

2) A “Two-state (or multistate) Brownian flip-flop” [32] is a<br />

machine that can reversibly change the distribution or<br />

position of a Brownian substrate between two (or more)<br />

distinguishable sites <strong>and</strong> can be reset without restoring the<br />

original distribution of the substrate (Figure 26c). The<br />

statistical distribution of the substrate cannot be determined<br />

from the state of the flip-flop but rather is<br />

determined by the history of operation of the machine.<br />

Rotaxane 85 is an example of a two-state Brownian flipflop.<br />

3) A “Brownian motor” [32,255] is a machine that can repetitively<br />

<strong>and</strong> progressively change the distribution or position<br />

of a Brownian substrate,during which the machine is reset<br />

without restoring the original distribution or position of<br />

the substrate (Figure 27). Like a flip-flop,a Brownian<br />

motor affects a system as a function of the pathway that<br />

the machine takes,not as a function of state. [6] Indeed,<br />

Figure 27. Schematic representations of two types of Brownian<br />

motors: a) a two-stroke rotary motor <strong>and</strong> b) a three-compartment<br />

translational motor or pump. [32]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

even at the machine s steady state it can produce a net flux<br />

of a substrate in a given direction. The rotary motors<br />

designed by Feringa <strong>and</strong> co-workers (see Section 2.2) are<br />

examples of this category of machines. A catenane-based<br />

two-stroke rotary Brownian motor in which the substrate<br />

is repetitively transported between two sites through two<br />

alternating pathways is given in Section 4.6.3.<br />

4) In addition to switches,flip-flops,<strong>and</strong> motors,other<br />

machines can be envisaged that exploit selective,controlled<br />

manipulation of Brownian moieties (parts of the<br />

machine molecule or a substrate) far from equilibrium. By<br />

combining balance-breaking <strong>and</strong> linking/unlinking steps<br />

with Boolean logic functions,machines that can move a<br />

substrate (or itself) in different directions (Figure 28a) or<br />

sort <strong>and</strong> separate different ions (Figure 28b) can be<br />

invented. We have little doubt that many other types of<br />

machines that exploit <strong>and</strong> control non-equilibrium distributions<br />

of conformations <strong>and</strong> co-conformations will be<br />

designed <strong>and</strong> realized in the future.<br />

The requirements for linking <strong>and</strong> unlinking,for example,<br />

can also be achieved by using a substrate that has two or more<br />

sites that can interact with the track. In its simplest form,this<br />

requires a “walker” with at least two “feet” (similar to<br />

biological molecular motors such as kinesin) <strong>and</strong> directional<br />

motion along the track can then occur by several different<br />

categories of energy ratchet-based mechanisms,for example,<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

119


Reviews<br />

Figure 28. Schematic representations of some compartmentalized<br />

molecular-level machines that combine ratcheting <strong>and</strong> escapement<br />

with Boolean logic operations: a) a four-compartment Brownian<br />

machine that pumps a substrate in a given (variable) direction.<br />

Deciding which one of the purple or yellow gates is used for ratcheting<br />

determines whether the substrate is transported to the top compartment<br />

or the bottom. b) A four-compartment Brownian machine that<br />

sorts <strong>and</strong> separates different ions (for example, red = Na + , blue = K + ).<br />

A logic operation providing selective access through each of the purple<br />

<strong>and</strong> yellow gates ensures one type of ion is pumped into each<br />

compartment. [32]<br />

“passing-leg” (Figure 29 a) <strong>and</strong> “inchworm” (Figure 29 b).<br />

Using the ideas outlined above (<strong>and</strong> in Section 1.4.3) it<br />

appears to us that each of these mechanisms can be realized<br />

through several different permutations of binding sites on the<br />

track (which may be active or passive) <strong>and</strong> on the walker<br />

“feet” (which may also be active or passive),but the key<br />

elements of the mechanism are the same. Compartmentalization<br />

is achieved by having,at all times,at least one foot<br />

kinetically associated with a particular site on the track,while<br />

a free foot seeks out the most preferable binding site for itself<br />

by Brownian motion. Directionality is achieved by the<br />

relative arrangement of stations in the region of space that<br />

the free foot can explore—controlled by the mode of walking<br />

(stepping or inchworm) <strong>and</strong> the length of linker between the<br />

feet (which could also be varied in some mechanisms). In<br />

Section 4.6.2 we will see that an inchworm-type mechanism<br />

around a cyclic track can be achieved in a [3]catenane where<br />

the two “walking” rings are not joined,although it would not<br />

be possible to extend this to a repetitive linear track without<br />

connecting them together. Such systems which operate solely<br />

through changes in the walker itself may be regarded as “selfpropelled”<br />

walkers (see Section 6.2).<br />

Just like compartmentalized molecular machines in which<br />

linking/unlinking is controlled by changes in the track<br />

(Figures 26–28),one can also envisage walkers which incorporate<br />

more complex functions,such as Boolean logic<br />

operations to select between a choice of pathways (Figure<br />

29c). Furthermore,an information ratchet mechanism<br />

may also be used to control the directionality of a walker,for<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 29. Schematic representations of compartmentalized molecular<br />

machines which use attachment of the substrate to a track to prevent<br />

a Brownian substrate from adopting a statistical thermodynamic<br />

equilibrium distribution. a) A “passing-leg” walker is ratcheted by<br />

fixing the “front” foot to the track while the rearmost foot swings<br />

forward to bind at a site further down the track in the direction of<br />

travel. b) In an “inchworm” walker, the order of the feet is prevented<br />

from changing (for example, by using rings to mechanically link to the<br />

track as illustrated here) <strong>and</strong> ratcheting occurs by first fixing the rear<br />

foot to the track while the front moves, then fixing this foot in place<br />

while the rear one moves. c) These mechanisms can be combined with<br />

more complex functions such as Boolean logic operations to allow a<br />

passing-leg walker to choose between different pathways. d) Information<br />

ratchet mechanisms can also be employed by walker compartmentalized<br />

machines, such as a “burnt-bridges” walker which destroys<br />

the track to the rear whenever a step is taken in the intended direction<br />

of travel.<br />

example a walker that irreversibly destroys its track when<br />

making a step in one direction,but not the other (Figure 29 d).<br />

4.5. Controlling Rotational Motion in Rotaxanes<br />

Macrocycle pirouetting in rotaxanes presents two challenges<br />

in terms of control: the frequency of the r<strong>and</strong>om<br />

motion <strong>and</strong> its directionality. The former can be achieved<br />

through temperature,structure,electric fields,light,<strong>and</strong><br />

solvent effects; the latter has yet to be demonstrated. [256]<br />

Alternating current (a.c.) electric fields represent an ideal<br />

stimulus with which to control submolecular dynamics in<br />

many applications. In the two [2]rotaxanes 86 <strong>and</strong> (E)-87<br />

(Scheme 50),it was observed that application of a.c. fields of<br />

around 50 Hz resulted in unusual Kerr effect responses which<br />

are unique to the interlocked architecture (they are not<br />

observed for either of the components alone). [257] Increasing<br />

120 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 50. [2]Rotaxanes 86 <strong>and</strong> 87 in which the rate of pirouetting<br />

can be controlled by application of an alternating current electric<br />

field. [257]<br />

the temperature had the same effect as decreasing the field<br />

strength,namely,enhancement of the response. The same sort<br />

of Kerr effects have previously been seen with rigid-rod<br />

polymers <strong>and</strong> correlated with the dynamics of the polymer<br />

backbones. [258] Variable temperature (VT) 1 H NMR experiments<br />

<strong>and</strong> molecular-modeling studies confirmed that macrocycle<br />

pirouetting is the only dynamic process in the rotaxanes<br />

on this timescale. The experimental <strong>and</strong> theoretical studies<br />

also predict the slightly more complex Kerr effect response<br />

observed experimentally for 87 relative to that of 86. It was<br />

concluded that application of an a.c. electric field<br />

attenuates macrocycle pirouetting in 86 <strong>and</strong> 87,<br />

probably by polarizing <strong>and</strong> strengthening intercomponent<br />

hydrogen bonding. The extent of the<br />

dampening can be varied with the strength of the<br />

applied field; even modest fields of about 1 V cm 1<br />

produce pirouetting rate reductions of two to three<br />

orders of magnitude.<br />

An alternative strategy for affecting pirouetting<br />

rates is to apply a stimulus which directly<br />

alters the structure of the thread or macrocycle.<br />

This has been demonstrated to great effect for<br />

olefin-containing [2]rotaxanes (E)/(Z)-87–89<br />

(Scheme 51). [259] The decrease in intercomponent<br />

binding affinity on photoisomerization of the<br />

fumaramide units in (E)-87–89 to the cis-maleamide<br />

isomers gives a huge increase in the rate of<br />

pirouetting of more than six orders of magnitude.<br />

The switching process is reversible: subjecting the<br />

maleamide rotaxanes to heat or a suitable catalyst results in<br />

reisomerization to the more thermally stable trans-olefin<br />

isomers,with accompanying reinstatement of the strong<br />

hydrogen-bonding network.<br />

Binding of sodium or potassium cations to a crown ether<br />

embedded in a rotaxane macrocycle has been shown experimentally<br />

to retard co-conformational Brownian motion,<br />

particularly pirouetting. [260] Removal of the metal template<br />

from a rotaxane can induce rotation of the ring relative to the<br />

thread, [261] while incorporating two different coordination<br />

sites into a macrocycle can be used to control the rotational<br />

orientation of components in a manner analogous to the<br />

translational shuttling modes discussed in Section 4.3.4. [262]<br />

This pirouetting motion,however,tends to be faster than the<br />

Scheme 51. Photoisomerization of [2]rotaxanes (E)-87–89 which results<br />

in pirouetting rate enhancements of up to six orders of magnitude. [259]<br />

The reverse process (Z!E isomerization) can be effected by heating a<br />

0.02m solution of the Z rotaxanes at 400 K, generating bromine<br />

radicals (cat. Br 2, hn 400 nm), or through reversible Michael addition<br />

of piperidine (RT, 1 h).<br />

analogous shuttling modes <strong>and</strong> in the particular example<br />

illustrated in Scheme 52 for [Cu(90)],the pirouetting rate is<br />

further enhanced by using an unhindered bipyridine lig<strong>and</strong> in<br />

the thread,from which the bulky stoppers are well separated.<br />

[262e]<br />

Scheme 52. Electrochemically triggered rapid pirouetting motion in metal-templated<br />

[2]rotaxane [Cu(90)]. [262e]<br />

4.6. Controlling Rotational Motion in Catenanes<br />

4.6.1. Two-Way <strong>and</strong> Three-Way Catenane Positional Switches<br />

Angew<strong>and</strong>te<br />

Chemie<br />

As with rotaxanes,stimuli-induced structural changes can<br />

be used to alter the rates of the thermal intercomponent<br />

motions in catenanes. For example,photoisomerization of an<br />

azobenzene unit in a p-donor/acceptor [2]catenane was found<br />

to influence the rate of dynamic processes,presumably by<br />

altering the size of the macrocycle cavity. [263] Electrochemical<br />

reduction of a homocircuit benzylic amide [2]catenane<br />

completely halts the circumrotational process as a result of<br />

the formation of an intercomponent covalent bond between<br />

the macrocycles. [264] Protonation of a porphyrin free base in a<br />

catenane was found to result in a rearrangement to minimize<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

121


Reviews<br />

repulsive electrostatic interactions,thus altering the rate of<br />

pirouetting of a tetracationic cyclophane. [191w,265] Electrochemical<br />

studies suggest similar conformational <strong>and</strong> dynamic<br />

changes can occur on reduction of the cyclophane. [266]<br />

The fundamental principles for controlling the position of<br />

a macrocycle on a thread in a rotaxane <strong>and</strong> the relative<br />

positions <strong>and</strong> orientations of the rings in a catenane are the<br />

same. For example,the behavior of amphiphilic homocircuit<br />

[2]catenane 91 is governed by the same driving forces that<br />

cause solvent-induced shuttling in amphiphilic shuttle 58 (see<br />

Sections 4.2.1 <strong>and</strong> 4.3.10). [267] In halogenated solvents such as<br />

CDCl 3,the two macrocycles of 91 interact through hydrogen<br />

bonds (amido-91,Scheme 53,also the structure observed in<br />

Scheme 53. Translational isomerism in an amphiphilic benzylic amide [2]catenane 91. [267]<br />

the solid state) whereby each constitutionally identical ring<br />

adopts a different conformation,one effectively acting as the<br />

host (convergent H-bonding sites) <strong>and</strong> the other the guest<br />

(divergent H-bonding sites). Pirouetting of the two rings<br />

interconverts the host–guest relationship rapidly on the NMR<br />

timescale in CDCl 3 at RT. In a hydrogen-bond-disrupting<br />

solvent such as [D6]DMSO,however,the preferred coconformation<br />

has the amide groups exposed on the surface<br />

where they can interact with the surrounding medium,while<br />

the hydrophobic alkyl chains are buried in the middle of the<br />

molecule to avoid disrupting the structure of the polar solvent<br />

(alkyl-91,Scheme 53). [268]<br />

In heterocircuit [2]catenanes such as 92 4+ the change in<br />

the position of one macrocycle with respect to the other is<br />

even more reminiscent of shuttling in [2]rotaxanes (Scheme<br />

54). [184l,228a] In the ground state (92 4+ ) the tetracationic cyclophane<br />

(blue) mostly sits over the more electron rich TTF<br />

station (green). Oxidation of the TTF (green!pink) to either<br />

its radical cation or dication (92 5+ or 92 6+ ) can be achieved<br />

chemically or electrochemically <strong>and</strong> results<br />

in the cyclophane being positioned over the<br />

dihydroxynaphthalene unit (DNP,red). The<br />

movement can be reversed by reduction of<br />

the TTF unit back to its neutral state. Just as<br />

for the rotaxane derivatives in this series<br />

(Section 4.3.4),[2]catenane 92 4+ exhibits<br />

rapid shuttling away from the TTF station<br />

<strong>and</strong> slower kinetics for the return<br />

step. [228f,269] There is,of course,no control<br />

over the direction in which the motion<br />

occurs; the cyclophane has a choice of two<br />

routes between the stations in both directions<br />

<strong>and</strong> half the molecules will change<br />

position in one direction <strong>and</strong> the other half in the other. [270,271]<br />

Electrochemical co-conformational switching has also<br />

been observed for [2]catenane 93,which features a crown<br />

ether threaded onto a large macrocycle containing two cyclen<br />

rings containing different metal ions (Scheme 55). [272] In both<br />

the solution <strong>and</strong> solid states,the crown ether sits over the<br />

Scheme 54. Chemically <strong>and</strong>/or electrochemically driven translational isomer switching of [2]catenane 92 4+ to 92 5+ /92 6+ [184l, 228a]<br />

.<br />

Scheme 55. Electrochemically induced co-conformational changes in a heterodinuclear [2]catenane. [272]<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

122 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Ni II center. However,the Cu II ion is more easily oxidized than<br />

Ni II ion,<strong>and</strong> the resulting Cu III center provides an effective<br />

electron-poor station for the p-donor macrocycle. Subsequent<br />

oxidation of the nickel ion moves the crown ether back to its<br />

original site.<br />

In the classical metal-templated [2]catenates pioneered by<br />

the Sauvage group (for example,[Cu I (94)],Scheme 56), [273]<br />

Scheme 56. The original metal-templated [2]catenate, [Cu I (94)]. [273]<br />

Removal of the Cu I ion results in reorganization of the organic lig<strong>and</strong>s<br />

to present the diphenylphenanthroline units (light blue) on the outer<br />

surface. [274] The resultant [2]caten<strong>and</strong> free lig<strong>and</strong> can complex a wide<br />

range of metal ions. [275,276]<br />

removal of the metal template to give the corresponding<br />

[2]caten<strong>and</strong> often results in significant co-conformational<br />

changes. In nonpolar solvents <strong>and</strong> the solid state,the<br />

polyether chains are buried in the center of the molecule,<br />

Scheme 57. Reversible co-conformational switching in a hybrid coordination/charge transfer [2]catenane. [277]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

thus presenting the heterocyclic lig<strong>and</strong>s to the outside. [274] The<br />

process is often fully reversible—indeed,complexation of a<br />

variety of cations can be achieved with the [2]caten<strong>and</strong><br />

(including Li + ,Fe 2+ ,Co 2+ ,Ni 2+ ,Cu + ,Zn 2+ ,Ag + ,Cd 2+ ,<strong>and</strong><br />

even H + ),in each case restoring the original complexed coconformation<br />

shown in Scheme 56. [275,276]<br />

Introducing a second,orthogonal,set of possible interactions<br />

between the two rings can make the co-conformational<br />

changes better defined. The presence of p-electrondonor<br />

(red) <strong>and</strong> p-electron-acceptor (dark blue) units on the<br />

two macrocycles in [2]catenate [Cu I (95)] 5+ (Scheme 57) [277]<br />

means that a precisely defined co-conformation ensues on<br />

demetalation to give [2]caten<strong>and</strong> 95 4+ ,which is stabilized by<br />

p–p charge-transfer interactions. A similar switching process<br />

can also be achieved reversibly simply by protonation <strong>and</strong><br />

deprotonation (Scheme 57). Anion binding has been shown<br />

to result in an alteration of the co-conformational arrangement<br />

of a bipyrrole-based [2]catenane templated by hydrogen-bonding<br />

interactions. [278]<br />

In a series of [2]catenates formed around square-planar<br />

templates,demetalation does not result in a significant coconformational<br />

change. [279] [2]Catenane 96 was generated in<br />

this fashion,but it was then found that the templating metal<br />

(Pd II ) can either be reinserted along with concomitant<br />

deprotonation of two amide nitrogen atoms to give the<br />

original catenate or else complexation to only one ring can be<br />

achieved,thereby producing a half-turn in the relative<br />

orientation of the components. [280] All three forms are fully<br />

interconvertible (Scheme 58) <strong>and</strong> can be observed in both<br />

solution <strong>and</strong> the solid state.<br />

Steps towards photochemical switching in metal-templated<br />

catenates have also recently been taken,by making use<br />

of the ready access of dissociative d-d* excited states in<br />

ruthenium(II) complexes with distorted octahedral geome-<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

123


Reviews<br />

Scheme 58. Half-rotation in a [2]catenane through interconvertible Pd II coordination modes. [280]<br />

tries. [281] Irradiation (l > 300 nm) of [2]catenate 97 results in<br />

dissociation of the bipyridyl unit (red),which allows free<br />

movement of the two rings (Scheme 59). The vacant coordination<br />

sites on the metal ion are filled by chloride ions added<br />

to the reaction mixture (shown) or by a coordinating solvent<br />

such as acetonitrile (not shown). [282] The recomplexation<br />

reaction is achieved simply by heating. [283]<br />

Sauvage <strong>and</strong> co-workers have also demonstrated both<br />

electrochemical <strong>and</strong> photochemical control over ring motions<br />

in hetero-[2]catenate 98 (Scheme 60 a) which is closely related<br />

to molecular shuttle 74 (see Section 4.3.4). [284] The behavior<br />

observed for the catenate is essentially the same as its<br />

rotaxane analogue. The kinetics for the intercomponent<br />

motions are again slow relative to other molecular machines,<br />

although some differences are observed between the catenane<br />

<strong>and</strong> rotaxane on account of easier access of the solvent<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

or ionic species to the metal center<br />

in the rotaxane,thus stabilizing<br />

transition states on the way to<br />

species with higher coordination<br />

numbers.<br />

The related homocircuit [2]catenate<br />

[Cu(99)] (Scheme 60b),in<br />

which each ring contains a bidentate<br />

dpp unit <strong>and</strong> tridentate terpy site,<br />

exhibits more complicated behavior.<br />

[285]<br />

In dpp,dpp-[Cu I (99)],the<br />

copper ion coordinates to the two<br />

dpp units in the usual tetrahedral<br />

arrangement. Oxidation of the<br />

metal center (by either chemical or<br />

electrochemical means) reverses the<br />

order of preference for coordination<br />

numbers (the preferred order for<br />

Cu II is 6 > 5 > 4). The result is circumvolution<br />

of the rings to give the<br />

preferred hexacoordinated species<br />

terpy,terpy-[Cu II (99)]. It was demonstrated<br />

that this process occurs by<br />

revolution of one ring to give an<br />

intermediate five-coordinate species<br />

(dpp,terpy-[Cu II (99)]). [285]<br />

In<br />

comparison to many related transition-metal-coordinated<br />

systems,the process is relatively fast,with the lig<strong>and</strong><br />

rearrangement occurring on the timescale of tens of seconds.<br />

The process is completely reversible,via the same fivecoordinate<br />

geometry,on reduction to Cu I (that is,via<br />

dpp,terpy-[Cu I (99)]).<br />

Accordingly,this catenate adopts three different translational<br />

isomeric forms (six different states when the oxidation<br />

state of the metal ion is considered). The intermediate fivecoordinate<br />

states are only transient,however,<strong>and</strong> could not<br />

be isolated. A system which displays three distinct,stable<br />

states is [2]catenane 100 (Scheme 61). This catenane,<strong>and</strong> the<br />

related [3]catenane 101,extend the olefin photoisomerization<br />

strategy utilized in [2]rotaxanes 80, 83,<strong>and</strong> 87–89 to create the<br />

first examples of stimuli-driven sequential <strong>and</strong> unidirectional<br />

rotation in interlocked molecules,thereby addressing the key<br />

Scheme 59. Photoinduced selective decomplexation in a ruthenium(II) [2]catenate. [282] In the decomplexed form [97 2+ ·2 Cl ], no bonding<br />

interactions exist between the two rings <strong>and</strong> free rotation occurs. The co-conformation shown is purely illustrative <strong>and</strong> does not indicate a<br />

preferred arrangement.<br />

124 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 60. a) Heterocircuit switchable catenate [Cu(98)]. [284] b) Oxidation-state-controlled switching of [2]catenate [Cu(99)] between three distinct<br />

co-conformations. [285]<br />

Scheme 61. [2]Catenane 100 <strong>and</strong> [3]catenane 101, shown as their<br />

E,E isomers. [286]<br />

difference between shuttling in two-station rotaxanes <strong>and</strong><br />

rotation in catenanes (closely related to the difference<br />

between a switch <strong>and</strong> a motor,see Section 1.1): the issue of<br />

directionality. [286]<br />

Sequential movement of one macrocycle between three<br />

stations on a second ring requires independent switching of<br />

the affinities for two of the units so as to change the relative<br />

order of binding affinities,as shown schematically in<br />

Figure 30. [286] In 100 this is achieved (Scheme 61) by employing<br />

two fumaramide stations with differing macrocycle binding<br />

affinities (steric mismatching of some tertiary amide<br />

rotamers disfavor the methylated station B),one of which<br />

(station A,green) is located next to a benzophenone unit.<br />

This allows selective photosensitized isomerization of station<br />

A at 350 nm,before photoisomerization of the other<br />

fumaramide station (station B,red) at 254 nm. The third<br />

station (station C,orange),a succinic amide ester,is not<br />

photoactive <strong>and</strong> is intermediate in macrocycle binding affinity<br />

Figure 30. Stimuli-induced sequential movement of a macrocycle<br />

between three different binding sites in a [2]catenane. [286]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

125


Reviews<br />

between the two fumaramide stations <strong>and</strong> their maleamide<br />

counterparts. A fourth station,an isolated amide group<br />

(shown as D in (E,E)-101) which can make fewer intercomponent<br />

hydrogen bonding contacts than A,B,or C,is also<br />

present but only plays a significant role in the behavior of the<br />

[3]catenane.<br />

Consequently,in the initial state (state I,Figure 30),the<br />

small macrocycle resides on the green,nonmethylated<br />

fumaramide station of [2]catenane 100. Isomerization of this<br />

station (irradiation at 350 nm,green!blue) puts the system<br />

out of co-conformational equilibrium <strong>and</strong> the macrocycle<br />

changes its net position to the new energy minimum on the<br />

red station (state II). Subsequent photoisomerization of this<br />

station (irradiation at 254 nm,red!purple) displaces the<br />

macrocycle to the succinic amide ester unit (orange,state III).<br />

Finally,heating the catenane (or treating it with photogenerated<br />

bromine radicals or piperidine) results in isomerization<br />

of both the Z olefins back to their E forms (purple!<br />

red <strong>and</strong> blue!green) so that the original order of binding<br />

affinities is restored <strong>and</strong> the macrocycle returns to its original<br />

position on the green station (state I).<br />

The 1 H NMR spectra of each diastereomer show excellent<br />

positional integrity of the small macrocycle at all stages of the<br />

process,but the rotation is not directional—over the complete<br />

sequence of reactions,an equal number of macrocycles go<br />

from A,through B <strong>and</strong> C,back to A again in each direction.<br />

4.6.2. Directional Circumrotation: A [3]Catenane Rotary Motor<br />

To bias the direction the macrocycle takes from station to<br />

station in a catenane such as 100,kinetic barriers are required<br />

to restrict Brownian motion in one direction at each stage <strong>and</strong><br />

bias the path taken by the macrocycle. Such a situation is<br />

intrinsically present in [3]catenane 101 (Scheme 61). [286]<br />

Irradiation of (E,E)-101 at 350 nm causes counterclockwise<br />

(as drawn) rotation of the light blue macrocycle to the<br />

succinic amide ester (orange) station to give (Z,E)-101.<br />

Isomerization (254 nm) of the remaining fumaramide group<br />

causes the other (purple) macrocycle to relocate to the single<br />

amide (dark green) station ((Z,Z)-101) <strong>and</strong>,again,this occurs<br />

counterclockwise because the clockwise route is blocked by<br />

the other (light blue) macrocycle. This “follow-the-leader”<br />

process,in which each macrocycle in turn moves <strong>and</strong> then<br />

blocks a direction of passage for the other macrocycle,is<br />

repeated throughout the sequence of transformations shown<br />

in Figure 31. After three diastereomer interconversions,<br />

(E,E)-101 is again formed,but 3608 rotation of each of the<br />

small rings has not yet occurred,they have only swapped<br />

places. Complete unidirectional rotation of both small rings<br />

occurs only after the synthetic sequence (a)–(c) has been<br />

completed twice.<br />

The directionality of the movement of the small macrocycles<br />

in Figure 31 could only be inferred from dynamic<br />

studies on related rotaxane-based molecular shuttles. [286]<br />

Nevertheless,these imply that stimuli-induced rotation carried<br />

out at 195 K occurs with extremely high (> 99%)<br />

directional fidelity. The DG ° value for background rotation<br />

of the small rings in (E,E)-101 is greater than 23 kcalmol 1 ,<br />

which means that the half-life for r<strong>and</strong>om circumrotation at<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 31. Stimuli-induced unidirectional rotation in a four-station<br />

[3]catenane, 101. [286] Conditions: a) irradiation at 350 nm; b) irradiation<br />

at 254 nm; <strong>and</strong> c) D; or catalytic ethylenediamine, D; or catalytic Br 2,<br />

irradiation at 400–670 nm.<br />

195 K is over 1.5 years. This design could only be used to do a<br />

limited amount of progressive mechanical work since the only<br />

way that the ring movements are ratcheted is the relatively<br />

modest DG ° value for background rotation. The efficiency of<br />

this motor is either poor (< 1 % based on irradiated photons)<br />

or modest (ca. 17 %,based on molecules that unidirectionally<br />

rotate during one sequence of reactions). Unlike the directionally<br />

rotating systems introduced by the research groups of<br />

Kelly (Section 2.1.2) <strong>and</strong> Feringa (Sections 2.1.2 <strong>and</strong> 2.2),<br />

[3]catenane 101 is achiral.<br />

It seems likely that the designs (<strong>and</strong> modes of operation)<br />

of synthetic molecular machines will provide significant<br />

feedback for the underst<strong>and</strong>ing <strong>and</strong> development of new<br />

fluctuation-driven transport mechanisms. If we compare<br />

[2]catenane 100 with [3]catenane 101,for example,we can<br />

see that they share a common track or potential-energy<br />

surface that is manipulated by the same set of chemical<br />

reactions. However,in [3]catenane 101 the Brownian particle<br />

(macrocycle) is transported directionally while in [2]catenane<br />

100 it is not. The only difference between the two machines is<br />

the “concentration” of the Brownian particles on the<br />

potential-energy surface. It seems possible that such a<br />

mechanism could account for the dependence of some<br />

biological pumps on the substrate concentration.<br />

4.6.3. Selective Rotation in Either Direction: A [2]Catenane<br />

Reversible Rotary Motor<br />

Clearly,just like their rotaxane analogues,catenanes such<br />

as 101 can be viewed as one macrocycle (the machine) which<br />

126 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

provides a potential-energy surface over which one (or more)<br />

smaller ring(s) (the Brownian substrate(s)) can be transported.<br />

By considering a flashing ratchet mechanism (Section<br />

1.4.2),it is possible to design a [2]catenane which is able<br />

to directionally rotate the smaller ring about the larger one in<br />

either direction in response to a series of chemical reactions<br />

(shown schematically in Figure 32). [51,287]<br />

This concept was realized in chemical terms through the<br />

synthesis <strong>and</strong> operation of catenane fum-(E)-102<br />

(Scheme 62). [51] Net changes in the position or potential<br />

energy of the smaller ring were sequentially achieved by:<br />

a) photoisomerization to maleamide (!mal-(Z)-102);<br />

b) desilylation/resilylation (!succ-(Z)-102); c) reisomerization<br />

to fumaramide (!succ-(E)-102); <strong>and</strong> finally,d) detritylation/retritylation<br />

to regenerate fum-(E)-102,the whole<br />

reaction sequence producing a net clockwise (as drawn in<br />

Scheme 62) circumrotation of the small ring about the larger<br />

one. Exchanging the order of steps (b) <strong>and</strong> (d) produced an<br />

equivalent counterclockwise rotation of the small ring.<br />

The simplicity of 102,together with the minimalist nature<br />

of its design,allows insight into the fundamental role each<br />

part of the structure plays in the operation of the rotary<br />

machine. [2]Catenane 102 is easily understood as an example<br />

of a compartmentalized molecular machine (see Section 4.4).<br />

The various chemical transformations perform linking/<br />

unlinking reactions (silylation/desilylation<br />

<strong>and</strong> tritylation/detritylation)<br />

<strong>and</strong> balance-breaking reactions<br />

(olefin isomerization (E!Z <strong>and</strong><br />

Z!E)). As with other compartmentalized<br />

molecular machines,the balance-breaking<br />

reactions control the<br />

thermodynamics <strong>and</strong> impetus for net<br />

transport; the linking/unlinking reactions<br />

control the relative kinetics <strong>and</strong><br />

ability to exchange. Raising the<br />

kinetic barriers ratchets transportation<br />

(see Section 4.4),thus allowing<br />

the statistical balance of the small ring<br />

to be subsequently broken without<br />

reversing the preceding net transportation<br />

sequence. Lowering the kinetic<br />

barrier allows escapement (see Section<br />

4.4) of a ratcheted quantity of<br />

rings in a particular direction—the<br />

element missing from rotaxane 85,<br />

which prevents it being a molecular<br />

motor.<br />

To obtain 3608 rotation of the<br />

small ring about the large ring,the<br />

four sets of reactions must be applied<br />

in one of two sequences,each taking<br />

the form: first a balance-breaking<br />

reaction; then a linking/unlinking<br />

step; then the second balance-breaking<br />

reaction; <strong>and</strong> finally,the second<br />

linking/unlinking step. This is the<br />

same manipulation of the potentialenergy<br />

surface shown for the flashing<br />

Figure 32. a) Schematic illustration <strong>and</strong> b) potential-energy surface for<br />

the blue ring in a minimalist [2]catenane rotary molecular motor. [51]<br />

Scheme 62. Reversible [2]catenane rotary motor 102. [51]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

127


Reviews<br />

ratchet in Figure 7 (Section 1.4.2). The direction of net<br />

rotation is determined solely by the way the balance-breaking<br />

<strong>and</strong> linking/unlinking steps are paired: an input of information.<br />

The efficiency or yields of the reactions—or the position<br />

of the ring at any stage (even if the machine makes a<br />

“mistake”)—are immaterial to the direction in which net<br />

motion occurs,as long as the reactions continue to be applied<br />

in the same sequence. Changing the pairings rotates the small<br />

ring in the opposite direction.<br />

The analysis of this deceptively simple molecule,particularly<br />

the separation of the kinetic <strong>and</strong> thermodynamic<br />

requirements for detailed balance,provides experimental<br />

insight into how an energy input is essential for directional<br />

rotation of a submolecular fragment by Brownian motion.<br />

Even though no net energy is used to power the motion,there<br />

has to be some processing of chemical energy for net rotation<br />

to be directional over a statistically significant number of<br />

molecules,a requirement that is absent if the equivalent<br />

motion is nondirectional. The amount of energy conversion<br />

required to induce directionality has an intrinsic lower limit,<br />

which corresponds to the difference in the binding energy of<br />

the fumaramide <strong>and</strong> maleamide binding sites,the same value<br />

that determines the directional efficiency of rotation <strong>and</strong> the<br />

maximum amount of work the motor can theoretically<br />

perform in a single cycle.<br />

5. <strong>Molecular</strong>-Level Motion Driven by External Fields<br />

In most of the synthetic molecular machines described so<br />

far,control over motion arises from the selective restriction of<br />

Brownian fluctuations through the manipulation of chemical<br />

structure,composition,<strong>and</strong> electronics to influence steric <strong>and</strong><br />

noncovalent bonding interactions. Common to all these<br />

approaches is that the forces exerted are short range—steric<br />

interactions are only felt over van der Waals distances,while<br />

bonding interactions (hydrogen bonds,charge-transfer interactions,or<br />

metal–lig<strong>and</strong> coordination) also operate over very<br />

short distances—<strong>and</strong> it is almost always thermal energy that<br />

powers the large-amplitude motions. It is well known that the<br />

application of external fields can cause bulk motion or<br />

orientation of molecular species,the most important technological<br />

application being the electric-field-directed alignment<br />

of liquid crystals. In addition to their more well-known<br />

applications,electrophoresis, [44l–n,q,v] dielectrophoresis, [44a,d,f,h,t]<br />

<strong>and</strong> photon pressure [44b,c] have been shown to provide<br />

potentials in which the motion of microscopic particles <strong>and</strong><br />

macromolecules can be manipulated through Brownian<br />

ratchet mechanisms. An emerging field of research,however,<br />

aims to use the long-range forces that can be applied by<br />

external electric <strong>and</strong> electromagnetic fields to influence<br />

submolecular motions. The effect of alternating current<br />

electric fields on the rate of r<strong>and</strong>om pirouetting in rotaxanes,<br />

which has already been discussed (Section 4.5),is just one<br />

example.<br />

Electric-field-controlled conformational switches have<br />

been proposed to create switchable molecular junctions for<br />

use in molecular electronics. [288] Computational investigations<br />

of 103 chemisorbed between two electrodes (Scheme 63)<br />

Scheme 63. A proposed molecular rectifier based on field-driven conformational<br />

switching. [289]<br />

have suggested that an electric-field-induced conformational<br />

change (arising from interaction of the field with the dipolar<br />

rotator,red) could result in a significant change in the<br />

conductance of the tunnel junction,thus resulting in a<br />

“conformational molecular rectifier”. [289] From a molecular<br />

electronics perspective,in both single-molecule junctions <strong>and</strong><br />

STM-contacted molecules (see Section 7),the ability of the<br />

electric current to induce dynamic motion of a molecule after<br />

transient electronic excitation is becoming a matter of both<br />

theoretical <strong>and</strong> practical investigation. [290]<br />

In terms of using long-range fields to drive the motions of<br />

a collection of molecular machines,the main challenge which<br />

has been addressed is the generation of submolecular rotary<br />

motion. [1k] The 3608 rotation of any rotor involves passage<br />

over a torsional potential-energy surface,with its particular<br />

series of minima <strong>and</strong> maxima. The effect of an external field<br />

could be to: 1) promote the system to an excited state where<br />

the torsional potential-energy surface is different,or 2) interact<br />

with a permanent,or induced,molecular dipole so as to<br />

force orientation in a particular direction. The interaction<br />

between the field <strong>and</strong> the rotor provides the energy to<br />

surmount kinetic barriers <strong>and</strong> overcome energy dissipation<br />

<strong>and</strong> thermal r<strong>and</strong>omization,so the field strength may need to<br />

be relatively large. With these intended systems,r<strong>and</strong>om<br />

thermal noise <strong>and</strong> dissipative effects such as friction would<br />

need to be minimized to maximize efficiency (for the<br />

problems in doing so,see Section 1.2). However,the requirements<br />

for generating unidirectional motion are no different<br />

from those under the thermally driven regime: asymmetry in<br />

the potential-energy surface <strong>and</strong> an energy input which can<br />

drive the system away from equilibrium.<br />

This type of rotary device is likely to<br />

require orientation of the rotational axis<br />

along a fixed direction by using surfacemounted,solid-state,or<br />

liquid-crystal<br />

rotors. However,the consideration of<br />

small molecules in the gas phase as<br />

models has allowed high-level quantum<br />

dynamics simulations of such a process.<br />

Fujimura <strong>and</strong> co-workers have studied<br />

the internal rotation in the single enantiomeric<br />

forms of chiral aldehyde 104,<br />

driven by an intense linearly polarized<br />

infrared laser pulse on the picosecond<br />

timescale (Scheme 64). [291,292] The chiral-<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Scheme 64. One enantiomer<br />

of chiral<br />

molecule 104 in<br />

which electric-fielddriven<br />

rotation<br />

around the C CHO<br />

bond was studied by<br />

quantum <strong>and</strong> classical<br />

mechanical sim-<br />

[291,293, 295,326]<br />

ulations.<br />

128 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

ity of 104 means that it has an asymmetric potential-energy<br />

surface for rotation about the C CHO bond,while energy<br />

sufficient to surmount the highest potential-energy barrier is<br />

provided by interaction between the electric field of the laser<br />

pulse <strong>and</strong> the permanent dipole of the aldehyde. The process<br />

is reminiscent of a rocking Brownian ratchet (Section 1.4.2.2),<br />

although here thermal energy is not required to surmount the<br />

kinetic barriers. The system can also be modeled as an<br />

ensemble of r<strong>and</strong>omly oriented rotors,but in either case the<br />

result is unidirectional motion <strong>and</strong> the two enantiomers rotate<br />

with equal magnitudes but in opposite directions (within the<br />

molecular frame of reference).<br />

In the above calculations dissipative effects were ignored.<br />

Inertia therefore plays a significant role <strong>and</strong> rotation is seen to<br />

continue even after the driving field is removed. As discussed<br />

in Section 1.2.2,it is somewhat debatable how applicable such<br />

conditions are to practical situations in which molecular<br />

machines might operate. Fujimura <strong>and</strong> co-workers have<br />

subsequently taken energy dissipation into account,<strong>and</strong><br />

have shown that,as one might expect,rotation only occurs<br />

under these conditions while the field is applied. [293] It was<br />

also found that on reducing the field intensity so that the<br />

field–dipole interaction falls below the maximum kineticbarrier<br />

height,rotation occurs in the opposite direction,with<br />

barrier crossings mediated by quantum effects. Such behavior<br />

bears a striking resemblance to rocked quantum-tunneling<br />

ratchets designed to pump electrons. In such cases,the<br />

direction of electron transport has been shown (both experimentally<br />

<strong>and</strong> theoretically) to switch with rising temperature,<br />

depending on whether quantum tunneling though the kinetic<br />

barriers (low temperatures) or thermally activated barrier<br />

crossing (high temperatures) is the dominant process. [294] The<br />

quantum control of rotational direction in enantiomerically<br />

pure 104 was further elucidated by studies in which the time<strong>and</strong><br />

frequency-resolved spectra of the laser electric field were<br />

examined <strong>and</strong> the conditions for generating rotation in each<br />

direction were optimized. [295]<br />

A number of computational studies of idealized graphite<br />

nanotube dynamics have been undertaken [296] in structures<br />

reminiscent of some of the futuristic “hard” nanotechnology<br />

machine components envisioned by Drexler. [297] The interactions<br />

between concentric nanotubes (that is,in doublewalled<br />

nanotubes (DWNTs) <strong>and</strong> multiwalled nanotubes<br />

(MWNTs)) have received considerable attention with<br />

regard to their predicted low friction for relative sliding <strong>and</strong><br />

rotational motions. [298] So-called “molecular bearings” in<br />

which the inner tube can be rotated within an outer sleeve<br />

have been studied, [299] as have “nanodrills” in which the<br />

interaction between the tubes is such that a rotational torque<br />

on one tube is converted into translation relative to the<br />

other. [300] Also proposed is the possibility of a gear effect for<br />

two parallel single-walled nanotubes (SWNTs) functionalized<br />

with benzyne-derived “teeth” along their length. [301] Potential<br />

methods for powering motions in all these systems clearly<br />

need to be considered. Attaching opposite charges to the<br />

inner tube of a DWNT molecular bearing in a molecular<br />

dynamics study enabled the coupling of the energy from a<br />

linearly polarized laser field with rotational modes of the<br />

system to be simulated. [302] A similar approach was used to<br />

study the theoretical motion of a SWNT gear system. [303]<br />

When the simulated gearing effect was used to transmit<br />

angular momentum to a noncharged nanotube,periods of<br />

induced directional rotation grew longer. [304] A further class of<br />

proposed nanotube-based mechanical devices are oscillators,<br />

in which the inner tube should be able to shuttle back <strong>and</strong><br />

forth inside an open-ended outer tube at gigahertz frequencies.<br />

[305] Again the problems of powering the proposed<br />

motions <strong>and</strong> providing environments in which they could be<br />

effective must be tackled. One scenario has been modeled in<br />

which the inner tube encapsulates potassium ions <strong>and</strong> its<br />

oscillatory motion is driven by an external electric field, [306]<br />

while another has proposed thermal expansion of gases to<br />

power the motion. [307] However,the arrival of the experimental<br />

techniques required to produce or operate any of<br />

these systems in practice appears to be distant. The chemical<br />

functionalization of nanotubes is,however,a rapidly developing<br />

field, [308] while improved manipulation techniques have<br />

allowed the telescoping of inner tubes from a MWNT,<strong>and</strong><br />

have demonstrated extremely low friction <strong>and</strong> a large driving<br />

force for the retraction of the extended structure. [309,310]<br />

Rotors driven by circularly or linearly polarized electric<br />

fields [311] have been investigated by Michl <strong>and</strong> co-workers.<br />

Their approach closely follows an experimental effort to<br />

create surface-mounted dipolar rotors <strong>and</strong> so their calculations<br />

had to tackle technologically relevant conditions <strong>and</strong><br />

complex structures. Dipolar chiral rhenium complex rotor<br />

105,was attached to a molecular grid constructed from<br />

[2]staffanedicarboxylate edges <strong>and</strong> dirhodium tetracarboxylate<br />

vertices (Scheme 65). [312,313] The theoretical rotational<br />

Scheme 65. Chiral dipolar rotor 105 which was studied computationally<br />

mounted on a 2D supramolecular framework <strong>and</strong> driven by a<br />

rotating electric field. [312]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

motion,driven by a circularly polarized electric field in a<br />

vacuum <strong>and</strong> at low temperature,was studied by molecular<br />

dynamics over a range of field frequencies <strong>and</strong> field strengths.<br />

At low field frequencies,the simulations suggest that the<br />

average lag angle (the position of the rotator in the rotating<br />

field coordinate system) depends only on the field strength,as<br />

the major restriction to unidirectional rotation is thermal<br />

energy. At higher frequencies,friction is the dominant<br />

hindrance to directional motion <strong>and</strong> appears to be linearly<br />

proportional to the driving frequency. Five different rota-<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

129


Reviews<br />

tional regimes were characterized,depending on the interplay<br />

of the field–dipole interaction,thermal energy,friction,<strong>and</strong><br />

intrinsic torsional barrier height. These range from synchronous<br />

motion,where the rotator slavishly follows the field,to<br />

situations where thermal energy or the torsional barrier<br />

dominates the behavior. [312,314]<br />

Michl <strong>and</strong> co-workers have synthesized nonpolar <strong>and</strong><br />

dipolar altitudinal rotors (that is,rotors with axles parallel to<br />

the surface) 106 <strong>and</strong> 107 (Scheme 66). [315] 19 F NMR spectroscopy<br />

showed that the barriers to rotation in 107 were<br />

extremely low in solution. Both systems have also been<br />

adsorbed on Au(111) <strong>and</strong> the surfaces studied by X-ray<br />

photoelectron spectroscopy (XPS),scanning tunneling<br />

microscopy (STM),<strong>and</strong> grazing incidence IR spectroscopy.<br />

It was found that,for a fraction of the molecules,the static<br />

electric field from the STM tip could induce an orientational<br />

change in the dipolar rotor but not the nonpolar analogue. [316]<br />

The propeller-like conformation of the tetraarylcyclobutadienes<br />

means that 107 can exist as three pairs (residual<br />

diastereomers) of helical enantiomers. Approximate molecular<br />

dynamics simulations have predicted an asymmetric<br />

rotational potential energy surface for at least one out of the<br />

three diastereomers,so that unidirectional rotation can be<br />

predicted for this isomer on application of an alternating<br />

current electric field. [317,318] The Michl research group has also<br />

reported initial synthetic investigations of an alternative,selfassembled,altitudinal<br />

rotor design,although no functional<br />

studies have yet been reported. [319] Jian <strong>and</strong> Tour have<br />

reported the synthesis of dipolar rotors such as 108 <strong>and</strong><br />

their monolayer formation on gold (Scheme 67),but have yet<br />

to communicate any results on its field-driven motion. [320]<br />

The above examples all consist of either permanent<br />

molecular dipole moments within a chiral structure,driven by<br />

achiral fields or else by a chiral field <strong>and</strong> therefore not<br />

requiring a chiral structure. Permanent molecular dipoles are<br />

not necessarily required however. If the rotator has a strongly<br />

anisotropic molecular polarizability,a dipole can be induced<br />

<strong>and</strong> rotated by an applied circularly polarized field. This has<br />

been proposed, [321] <strong>and</strong> experimentally achieved, [322] by subjecting<br />

chlorine molecules to an intense linearly polarized<br />

laser pulse,the polarization of which is rotating. In this<br />

“optical centrifuge” extremely high rotational energy levels<br />

can be accessed in very short time periods <strong>and</strong> the centrifugal<br />

force can induce bond dissociation.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Scheme 67. Dipolar rotor 108 designed to be attached to gold<br />

surfaces. [320]<br />

An approach which has yet to be investigated experimentally<br />

[323]<br />

is the transfer of momentum from a directional<br />

stream of atoms or molecules to bias the rotation of<br />

components of molecules fixed to a surface. Three possible<br />

arrangements can be envisaged: [1k] 1) fluid flow parallel to the<br />

rotor axle,combined with a chiral propeller-like rotator<br />

(windmill-like operation); 2) fluid flow perpendicular to the<br />

rotor axle,with faster flow on one side of the axis of rotation<br />

than the other (waterwheel-like operation); <strong>and</strong> 3) fluid flow<br />

perpendicular to the rotor axle,combined with rotator blades<br />

which interact with the fluid particles differently on either<br />

side (operating somewhat like an anemometer). Vacek <strong>and</strong><br />

Michl have modeled the second of these possibilities for chiral<br />

rotor 105 (Scheme 65) as well as an analogue bearing larger<br />

blades. [324] The molecular dynamics simulations,in which a<br />

supersonic jet of noble gas atoms was directed parallel to the<br />

rotor axis through the open grid structure on which it was<br />

mounted,suggested that high-density beams of lighter atoms<br />

(He or Ne,but not Ar or Xe) might indeed be able to induce<br />

Scheme 66. Nonpolar (106) <strong>and</strong> dipolar (107) altitudinal rotors which have been studied on Au(111) surfaces. Note that the rotator <strong>and</strong> flanking<br />

aryl rings are arbitrarily shown perpendicular to the surface for clarity. Interaction with the surface is through several atoms in the cyclopentadienyl<br />

“tentacle” substituents. [315,317]<br />

130 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

directional rotation,albeit at low temperatures <strong>and</strong> in<br />

competition with deformation of the rotor. [325]<br />

Fujimura <strong>and</strong> co-workers performed calculations that<br />

consider exploiting the different torsional potential-energy<br />

surfaces in the electronic ground <strong>and</strong> excited states of 104<br />

(Scheme 64) to achieve unidirectional rotation. [326] A femtosecond,UV-frequency<br />

pulse was used to excite the molecule<br />

to its first excited state. As a vertical transition does not<br />

correspond to a minimum in the excited-state potentialenergy<br />

surface,rotation is induced,the direction of which is<br />

governed by the gradient of the potential-energy surface in<br />

the Franck–Condon region (opposite in each enantiomer). In<br />

theory,if the molecules are returned to the electronic ground<br />

state by a second femtosecond pulse before the excited-state<br />

minimum is reached,then the angular momentum can be<br />

preserved <strong>and</strong> rotation continues in the same direction. Of<br />

course,Brownian thermal <strong>and</strong> frictional effects will very<br />

quickly degrade such motion <strong>and</strong> relaxation must next be<br />

taken into account to ascertain if sequential such “pump–<br />

dump” cycles would be able to maintain unidirectional<br />

rotation. A related approach was applied to theoretically<br />

consider rotation around the double bond in 109<br />

(Scheme 68). [327] In this case,an IR-frequency pulse would<br />

be used to excite vibrational motion in<br />

the deep torsional potential well of the<br />

double bond. In the first electronic<br />

excited state,the double-bond charac-<br />

Scheme 68. Axially<br />

chiral 1-(fluoromethylene)-4-methylcyclohexane<br />

(109), in<br />

which rotational<br />

motion around the<br />

double bond, initiated<br />

by an IR frequency followed<br />

by a UV-frequency<br />

laser pulse,<br />

has been studied<br />

computationally. [327]<br />

ter is significantly reduced <strong>and</strong> kinetic<br />

barriers to rotation are much smaller.<br />

On excitation to this state by using a<br />

UV laser,all the kinetic barriers could,<br />

in principle,be overcome <strong>and</strong> the<br />

instantaneous angular momentum at<br />

the moment of excitation preserved. A<br />

well-timed excitation pulse should<br />

therefore be able to select the direction<br />

in which rotation will continue in the<br />

excited state. Once again,if relaxation<br />

effects are taken into account,this<br />

rotation will soon be retarded <strong>and</strong> will<br />

certainly stop on return to the ground state. A timely return to<br />

a repulsive region on the ground-state surface,however,may<br />

be able to maintain the motion <strong>and</strong> allow a second cycle of<br />

excitation to be applied. The use of two electronic states of a<br />

double bond to generate unidirectional rotation can be<br />

related back to the directional motors from the Feringa<br />

research group,discussed in Section 2.2.<br />

As shall be discussed extensively in Section 8,the<br />

incorporation of molecular machines into ordered phases,<br />

such as at interfaces or in liquid crystals,is an important goal<br />

in current research. Crystalline solids,however,present a very<br />

different environment within which to explore submolecular<br />

motions. [328] The combination of crystalline order with<br />

addressable molecular motions,to yield so-called “amphidynamic”<br />

crystals,provides a novel <strong>and</strong> challenging situation for<br />

the operation of artificial molecular machines. To this end,<br />

Garcia-Garibay <strong>and</strong> co-workers have initiated a detailed<br />

investigation of the gas-phase <strong>and</strong> solution-phase dynamics as<br />

well as the solid-state packing <strong>and</strong> dynamics of a series of<br />

Angew<strong>and</strong>te<br />

Chemie<br />

molecules such as 110–113 (Scheme 69). [329,330] It is intended<br />

that the bulky trityl- (110, 112,or 113) or triptycyl-derived<br />

(111) portions (blue) should form the lattice framework <strong>and</strong><br />

remain static,while rotation of the phenylene (110, 111,or<br />

Scheme 69. <strong>Molecular</strong> “gyroscopes”. [329,330] A selection of rotor structures<br />

constructed with the intention of demonstrating controllable<br />

molecular-level motion in crystalline solids. The bulky components<br />

colored blue are intended to direct the formation of ordered crystalline<br />

arrays in which free volume around the rotator section (red) results in<br />

low barriers to rotation.<br />

113) or diamantane (112) rotator (red) is the single internal<br />

degree of freedom. The arrangement of a central rotator<br />

spinning within a framework which shields it from external<br />

contracts can be compared to a macroscopic gyroscope. [331]<br />

Key to the desired dynamics is the volume-conserving nature<br />

of the conformational process,together with the ability of the<br />

bulky framework to force packing arrangements which allow<br />

free volume around the rotator unit. The construction of<br />

analogues in which the central rotator contains a dipole is<br />

intended to create systems in which the motion can be<br />

controlled by an external electric field. [329d] The thermal solidstate<br />

dynamics of one such example, 113 (DG °<br />

13 kcal<br />

mol 1 for fluoroarene spinning),have been studied by both<br />

NMR <strong>and</strong> dielectric spectroscopies, [329j] while dynamic studies<br />

of the nonpolar analogues suggest that significantly lower<br />

barriers to rotation are obtainable on suitable functionaliza-<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

131


Reviews<br />

tion of the static framework, [329e,330a] or on incorporation of<br />

rotators with higher symmetry,as in 112. [329i,332]<br />

A series of fully encapsulated gyroscope structures have<br />

been prepared <strong>and</strong> studied in solution by Gladysz <strong>and</strong> coworkers.<br />

[333] The crystal structure of 114 indicates that<br />

sufficient free volume around the rotator should exist to<br />

permit a low barrier to rotation in the solid state (Scheme 70).<br />

Scheme 70. Fully encapsulated “gyroscope” structures 114 <strong>and</strong> 115. [333]<br />

Reduced-symmetry derivative 115,which also possesses a<br />

dipole moment as a possible h<strong>and</strong>le to externally control the<br />

motion,allowed the rotational motion to be observed in<br />

solution. However,the activation barrier for the process was<br />

found to be not insignificant (DG °<br />

11 kcalmol 1 ). [333] As<br />

was noted in Section 1.2.2,the effects of friction <strong>and</strong> thermal<br />

noise are extremely important on the molecular scale,but it is<br />

proposed that conditions may ultimately be found in which<br />

they are reduced to the extent that inertia in encapsulated or<br />

solid-state rotors becomes significant.<br />

Supramolecular crystalline structures offer a different<br />

approach to solid-state rotors. The crown ethers in the tripledecker<br />

supramolecular cation [Cs2([18]crown-6) 3] 2+ undergo<br />

rotational motion above 220 K in a crystalline environment<br />

with paramagnetic [Ni(dmit) 2] counterions (dmit 2 = 2thioxo-1,3-dithiole-4,5-dithiolate).<br />

[334]<br />

Furthermore,the<br />

motion is strongly coupled with the magnetic properties of<br />

the crystal,thus suggesting one potential control mechanism.<br />

It has already been demonstrated that the application of<br />

pressure to the crystal can retard or even stop the thermally<br />

driven motion. [334]<br />

6. Self-Propelled “Nanostructures”<br />

The self-propulsion of microscopic objects has fascinated<br />

scientists from a range of backgrounds for well over a century.<br />

Spontaneous physical phenomena such as “tears of wine” <strong>and</strong><br />

the motion of “camphor boats” engaged the minds of many<br />

eminent 19th century scientists,while the advent of microscopy<br />

revealed the deterministic motion of microorganisms to<br />

the biological community. Two distinct mechanisms—interfacial<br />

<strong>and</strong> mechanical—can be discerned in these natural<br />

phenomena <strong>and</strong> attention has recently turned to artificial<br />

devices,ultimately of molecular size,which can achieve the<br />

same results. [335] Just like the machines in Section 5,nanoscale<br />

self-propulsion devices are not powered by thermal energy.<br />

Field-driven mechanisms,however,address ensembles of<br />

molecules,whereas here we look at structures whose movements<br />

are driven by forces generated between each individual<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

machine <strong>and</strong> the surrounding medium or track. Such<br />

machines can either operate autonomously—where the<br />

energy for movement is constantly present—or non-autonomously—where<br />

the energy must be applied in a series of<br />

stimuli given in a particular order.<br />

6.1. Propulsion by Manipulating Surface Tension<br />

A difference in surface tensions on either side of a liquid<br />

droplet or gas bubble can cause its directional transport<br />

through a second liquid or over a solid substrate. [336] This<br />

phenomenon is known as the Marangoni effect <strong>and</strong> is<br />

responsible for a number of naturally observed spontaneous<br />

processes. [337] In artificial systems,Marangoni flows are<br />

usually associated with a temperature gradient [336a,338] <strong>and</strong><br />

have been proposed as a means of microfluidic pumping. [339]<br />

A similar effect can be achieved for a droplet sitting on a<br />

homogeneous surface if the droplet contains a species which<br />

adsorbs onto the surface. [340] Droplet motion is then driven by<br />

the irreversible modification of the surface free energy which<br />

affects the interfacial energy on either side of the droplet (see<br />

also Section 8.3.4)—a so-called “chemical” Marangoni<br />

effect. [341]<br />

Whitesides <strong>and</strong> co-workers have reported the autonomous<br />

movement of millimeter-scale metallic objects at the<br />

liquid/air interface. [342] On one side,the objects feature a small<br />

area of platinum metal which catalyzes the decomposition of<br />

hydrogen peroxide to water <strong>and</strong> dioxygen. In an aqueous<br />

solution of hydrogen peroxide,the metallic plates are<br />

propelled by the recoil force of the oxygen bubbles. Subsequently,Sen,Mallouk,Crespi,<strong>and</strong><br />

co-workers created<br />

smaller rod-shaped particles consisting of 1-mm-long platinum<br />

<strong>and</strong> gold segments. [343] These were suspended within an<br />

aqueous solution of hydrogen peroxide <strong>and</strong> again showed<br />

propulsion which is directly linked to the production of<br />

oxygen. In contrast to the system developed by Whitesides<br />

<strong>and</strong> co-workers,however,movement was found to be towards<br />

the oxygen-producing platinum region. It appears that the<br />

oxygen produced adheres to the hydrophobic gold surface in<br />

the form of nanobubbles,thus setting up a concentration<br />

gradient along the gold segment as the oxygen diffuses away<br />

from the platinum regions. It has been proposed that the<br />

resulting interfacial energy gradient causes motion of the rod<br />

towards the platinum end. However,the precise details of the<br />

propulsion mechanism are still a matter of investigation. [335,344]<br />

The driven <strong>and</strong> r<strong>and</strong>om thermal components of the motion<br />

could be distinguished at various viscosities for rods operating<br />

near an aqueous/organic interface. [345] The same technology<br />

has been used to create a rotational device, [346] while<br />

incorporation of a ferromagnetic metal into the nanorod<br />

design has allowed control over the directionality of movement<br />

by using an external magnetic field. [347] Independently,a<br />

nanorod system composed of nickel <strong>and</strong> gold regions has been<br />

developed by Ozin,Manners,<strong>and</strong> co-workers which undergoes<br />

both linear <strong>and</strong> rotational self-propelled motions in<br />

hydrogen peroxide (Figure 33),the latter seemingly a consequence<br />

of the tethering of one end of a rod to a surface<br />

impurity or defect. [348]<br />

132 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Figure 33. Optical microscope snapshots of a nickel/gold nanorod<br />

moving in a linear fashion (1!3) before becoming attached to a<br />

surface impurity <strong>and</strong> rotating counterclockwise (4!6). [348] Reprinted<br />

with permission from Ref. [348].<br />

All these devices are both autonomous<br />

<strong>and</strong> catalyst-based,which means<br />

that,unlike a camphor boat or a chemical<br />

Marangoni droplet,they do not have<br />

to carry a finite supply of “fuel” onboard.<br />

Clearly,however,they are not yet<br />

molecular in nature. Recently Feringa<br />

<strong>and</strong> co-workers tethered a synthetic<br />

dinuclear manganese catalase mimic to<br />

a silica microparticle <strong>and</strong> again observed<br />

motion powered by the decomposition<br />

of hydrogen peroxide. [349] Although<br />

functionalization of the microparticle is<br />

theoretically homogeneous,defects<br />

result in inhomogeneous bubble nucleation,thus<br />

providing the asymmetry that<br />

gives rise to directional powered movement.<br />

Even though the mechanism of<br />

propulsion in this case is not yet clear,it<br />

demonstrates that the catalytic domain<br />

of a self-propelled motor can be scaled<br />

down to a molecular size. It remains to<br />

be seen if there is a lower limit to the<br />

particle size for which these types of<br />

strategies are practical (in all current<br />

systems,the movement of particles less<br />

than a micrometer in length is indistinguishable<br />

from Brownwin motion),what<br />

fidelity of directionality is possible (particularly<br />

for small particles where Brownian<br />

motion becomes significant),<strong>and</strong><br />

whether or not it is possible to constrain the medium to a<br />

track along which the particle moves (for example,by<br />

confining the fluid to a microscopic capillary).<br />

6.2. <strong>Mechanical</strong> Self-Propulsion—Molecules that Can Swim?<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Most microorganisms use a different mechanistic<br />

approach to self-propulsion which is more akin to the<br />

mechanical swimming of an animal or propulsion of a<br />

macroscopic boat. However,the low Reynolds number at<br />

small length scales (see Section 1.2.2) puts constraints on the<br />

nature of productive swimming mechanisms. [350] This was<br />

probably first considered from a physical perspective by<br />

Purcell,who deduced that any useful mechanical mechanism<br />

must involve a nonreciprocal motion which breaks the timereversal<br />

symmetry. [27] Since the mechanism needs to be<br />

repetitive,this effectively corresponds to the requirements<br />

of a molecular-level motor in which the substrate is the<br />

components of the machine itself (Sections 1.4.2 <strong>and</strong> 4.4).<br />

Purcell proposed a simple way in which this could be<br />

achieved: utilization of a “swimmer” composed of three<br />

rigid parts,linked by two hinges (Figure 34 a). This device has<br />

recently been examined quantitatively [351] <strong>and</strong> reduced to a<br />

simpler one-dimensional problem by replacement of the<br />

hinges by simple springs (Figure 34b). [352] A number of other<br />

Figure 34. a) Nonreciprocal sequence of configurations for Purcell’s three-link swimmer, which could<br />

result in propulsion at low Reynolds number. [27,351] b) A one-dimensional relative of Purcell’s three-link<br />

swimmer. [352] c) Possible nonreciprocal motion in a three-part, two-hinge molecule. d) Possible nonreciprocal<br />

motion in a two-part, one-hinge molecule: cis–trans isomerization of an imine. [356]<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

133


Reviews<br />

swimming mechanisms involving nonreciprocal motion have<br />

also been investigated theoretically from the dual perspectives<br />

of modeling biological phenomena <strong>and</strong> designing microscopic<br />

devices. [353] Somewhat surprisingly (<strong>and</strong> perhaps analogous<br />

to the lack of molecular designs based on fluctuationdriven<br />

transport mechanisms),despite a recent success in<br />

creating a microscopic swimmer which operates by nonreciprocal<br />

motion, [354] there appears to have as yet been no<br />

investigation of such mechanisms at the molecular level.<br />

Incorporating two orthogonally addressable switches into a<br />

molecule would give a Purcell-like three-part-two-hinge/<br />

spring structure (Figure 34 c). Adding bulk,the cyclodextrin<br />

in a known rotaxane system, [355] for example,would increase<br />

the net change in the position of the mass. However,the<br />

requirements of stimuli-induced nonreciprocal nanoscale<br />

motion for a low Reynolds number swimmer could also<br />

potentially be fulfilled by a two-part-one-hinge/spring molecule,by<br />

using a chemical reaction that can be reversed by a<br />

different nuclear pathway to the original transformation; for<br />

example,the photoisomerization <strong>and</strong> thermal reisomerization<br />

of certain imines [356] (Figure 34d). Other mechanical molecular<br />

switches discussed elsewhere in this Review are plausible<br />

c<strong>and</strong>idates for linear mechanisms of the type shown in<br />

Figure 34 b. The major problem with the realization of any<br />

of these systems in a working molecular form will be carrying<br />

out the stimuli-fueled motions fast <strong>and</strong> frequently enough so<br />

that they are not swamped by the background Brownian<br />

motion.<br />

Repetitive directional rotation of any chiral object is,of<br />

course,also a nonreciprocal motion. Field-driven unidirectional<br />

rotation of a chiral molecular rotor (Section 5) which is<br />

not confined to a surface through any stator unit could result<br />

in its propulsion through a liquid in direct analogy to a screw<br />

propeller on a boat or the bacterial flagellar system. This<br />

situation has been studied theoretically with a view to using<br />

circularly polarized microwaves to spatially resolve a racemic<br />

mixture of “propeller-shaped” molecules. [357] It has been<br />

postulated that anchoring such rotors at a fixed location<br />

within a fluid could theoretically create a microfluidic<br />

pumping system. [312]<br />

7. <strong>Molecular</strong>-Level Motion Driven by Atomically<br />

Sharp Tools<br />

The development of STM [358] <strong>and</strong> AFM [359] has dramatically<br />

shifted the goalposts for what it is possible to see,<strong>and</strong><br />

potentially do,with synthetic molecular machines. For the<br />

first time these instruments offer scientists the opportunity to<br />

both observe <strong>and</strong> manipulate atomic <strong>and</strong> molecular events<br />

that operate not on ensemble-averaged populations of species<br />

but on single chemical entities. [360,361]<br />

The early STM manipulations of single atoms <strong>and</strong><br />

molecules had to be carried out at extremely low temperatures<br />

to suppress thermal motion. In one famous example,a<br />

single-atom electronic switch was created by using STM,with<br />

its operation based on changing the position of a xenon atom<br />

relative to two electrodes. [362] The manipulation of organic<br />

molecules requires careful balancing of adsorbant–adsorbate<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

interactions: molecule–substrate binding must be strong<br />

enough to prevent thermal motion at the temperature used,<br />

but not so strong that rupture of covalent bonds results<br />

instead of translation across the surface. [360e,f] These criteria<br />

were first met at room temperature in 1996 with porphyrin<br />

116 (Figure 35). [363] In the gas phase,<strong>and</strong> on a Cu(100) surface,<br />

Figure 35. a) Porphyrin 116 for which positioning was induced at room<br />

temperature on a Cu(100) surface by the STM tip. [363,364, 366] b) Decacyclene<br />

117, for which STM-tip-switched, thermally driven rotations are<br />

observed on a Cu(100) surface at coverages just below one monolayer<br />

(c <strong>and</strong> d). [374] Both molecules adsorb on the surface through the<br />

appended tert-butyl groups, with the main planar skeletons in the<br />

plane of the substrate. c) <strong>and</strong> d) Constant-current STM images of a<br />

clean Cu(100) surface with a submonolayer coverage of 117. Stationary<br />

molecules appear as a circular arrangement of six lobes (arising from<br />

the six tert-butyl groups) <strong>and</strong> gaps in the monolayer appear as dark<br />

regions. In (c), the marked molecule is at a gap in the monolayer but<br />

it is aligned with the lattice <strong>and</strong> therefore stationary. In (d), the same<br />

molecule has been moved a distance of 0.25 nm by the STM tip <strong>and</strong> is<br />

no longer bound tightly by its neighbors <strong>and</strong> it therefore rotates, which<br />

is observed as a blurring of the lobed structure. The STM images are<br />

reprinted with permission from Ref. [374].<br />

the di-tert-butylphenylene substituents are oriented perpendicular<br />

to the porphyrin unit <strong>and</strong> act as relatively sticky “legs”<br />

which separate the polyaromatic core from the surface,<br />

thereby fixing the molecules securely in place while modulating<br />

the extremely strong interaction of the planar central core<br />

with the substrate. [364] Pushing with an STM tip causes<br />

translational motion,which is facilitated by accompanying<br />

r<strong>and</strong>om rotations or “chattering” of the phenyl rings,which<br />

lowers the barrier to movement. [363,365,366]<br />

Interlocked architectures provide another means through<br />

which the correct balance of forces can be achieved,<br />

ultimately enabling submolecular co-conformational changes<br />

to be made in a single molecule. An early report on STM<br />

imaging of a benzylic amide [2]catenane deposited from<br />

solution onto highly oriented pyrolitic graphite (HOPG)<br />

showed rather weak adsorption on the surface. [367] A much<br />

134 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

larger [2]catenane was subsequently shown to form a tightly<br />

packed polycrystalline arrangement on the same surface,with<br />

the larger number of aromatic functionalities <strong>and</strong> greater<br />

flexibility leading to stronger adsorption. [368] Co-conformational<br />

motion in an interlocked architecture was achieved<br />

when an STM tip was used to reversibly move one or two<br />

adjacent cyclodextrin rings along the polyethylene glycol<br />

derived thread in a polyrotaxane adsorbed on MoS 2<br />

(Figure 36). [369] This abacus-like motion is reminiscent of an<br />

earlier report on the positional manipulation of C 60 molecules<br />

on Cu(111) surfaces,where the thermal motion of these rather<br />

poorly adsorbed species is restricted by confinement to a<br />

monatomic step edge. [370]<br />

Figure 36. Translocation of a cyclodextrin ring in a polyrotaxane<br />

induced by a scanning probe. [369] The ring (indicated by a yellow arrow)<br />

is moved first towards the left (a!b) then back to the right (b!c).<br />

The white box on image (a) indicates the region of detail which is<br />

magnified in (b) <strong>and</strong> (c). Reprinted with permission from Ref. [369].<br />

An STM tip has been used to apply a localized voltage to a<br />

thin film of bistable [2]rotaxanes which has resulted in<br />

conductance switching. [371] Based on the behavior of similar<br />

molecules in other solid-state experiments (see Section 8.3.1)<br />

it is proposed this effect is due to macrocycle shuttling<br />

triggered by tip-induced oxidation of the thread. A similar<br />

single-molecule switching phenomenon has been observed for<br />

an oligopeptide organized in a self-assembled monolayer. The<br />

peptide adopts one of two helical conformations depending<br />

on the polarization of the STM tip <strong>and</strong> the two states are<br />

characterized by different molecular heights <strong>and</strong> conductances.<br />

[372,373]<br />

STM-induced single-molecule positional switching has<br />

been employed to reversibly control thermally driven motions<br />

on surfaces. [374] At monolayer coverage on Cu(100),decacyclene<br />

117 forms a two-dimensional van der Waals crystal,<br />

while at coverages significantly below one monolayer,<br />

r<strong>and</strong>om thermal motions are extremely fast <strong>and</strong> not observable<br />

by STM. When coverage is a little less than a full<br />

monolayer,however,the 2D lattice has small gaps or “nanocavities”<br />

(Figure 35c,d). The molecules at the borders of these<br />

free spaces can sit in one of two positions: a highly symmetrical<br />

one,following the order of the adjacent lattice,or a<br />

less symmetrical one. Movement between these two sites can<br />

be effected by the STM tip. In the less symmetrical site,the<br />

molecule is still constrained by its neighbors,but is disengaged<br />

from some of the intermolecular interactions,so that it freely<br />

rotates in a r<strong>and</strong>om fashion at high speeds (Figure 35 d). [374,375]<br />

It has been noted that in the “engaged” state the potentialenergy<br />

profile for rotation is asymmetric,thus suggesting that<br />

such a system could be a starting point for a unidirectional<br />

Angew<strong>and</strong>te<br />

Chemie<br />

rotor. To date,however,there have been no further reports on<br />

the incorporation of the remaining features necessary for such<br />

a system. [376]<br />

While STM permits the manipulation <strong>and</strong> observation of<br />

complex molecular species adsorbed on conducting surfaces,<br />

AFM can be used to image nonconducting surfaces <strong>and</strong> can<br />

operate under ambient conditions,even on samples in<br />

solution. The use of an AFM cantilever as a force sensor or<br />

applicator (so-called “force spectroscopy”) has facilitated<br />

some remarkable single-molecule investigations of receptor–<br />

lig<strong>and</strong> interactions,of the entropic <strong>and</strong> enthalpic factors<br />

involved in biopolymer folding,<strong>and</strong> of the mechanical<br />

elasticity of biopolymers. [360j–p] <strong>Synthetic</strong> polymers are also<br />

amenable to this approach,although nonspecific forces at<br />

small tip–substrate separations <strong>and</strong> the physical size of<br />

available tips have impeded the application of AFM to<br />

small-molecule perturbations <strong>and</strong> measurements to<br />

date. [360g,i,377] It has recently been discovered that an AFM<br />

tip can be used to initiate the formation of regular arrays of<br />

deformations in thin films of simple rotaxanes (Figure 37). [378]<br />

The effect is unique to films made from interlocked molecules<br />

(it does not happen to films of the non-interlocked components,for<br />

example) <strong>and</strong> is a result of coupled nucleation<br />

recrystallization being favored by the ease of intercomponent<br />

mobility for these molecules. The features of the nanoscale<br />

dot arrays are easily varied by changing the thickness of the<br />

film,<strong>and</strong> show potential for application in information<br />

storage technology.<br />

Figure 37. a) Array of dots fabricated by individual line scans of the<br />

AFM tip on a 5-nm-thick film of a benzylic amide [2]rotaxane 89<br />

(Scheme 51) deposited on HOPG. [378] b) For a given thickness (here<br />

20 nm), the number of dots is linearly proportional to the scan length.<br />

The film thickness controls the characteristic size. c) A pattern made<br />

up of 31 lines with 45 dots each on an approximately 30 ” 30 mm 2 area<br />

of a thicker film. d) Proof-of-concept for information storage. The<br />

sequence “ec7a8” in the hexadecimal base corresponds to the<br />

number 968616.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

135


Reviews<br />

Despite the advances in the observation <strong>and</strong> manipulation<br />

of single-molecule dynamics,it is only recently that potential<br />

structural precursors for molecular machines specifically<br />

designed to be actuated at the single-molecule level by<br />

probe techniques have been investigated. Variations in<br />

conductance on deformation by atomically sharp probes<br />

have been reported for a number of molecular species,in<br />

particular fullerenes. [360s] For example,the angle between the<br />

porphyrin core <strong>and</strong> an individual di-tert-butylphenyl leg in 116<br />

can be reversibly manipulated by either STM or AFM,<br />

thereby leading to an order of magnitude switching of the<br />

STM tunneling current between the tip <strong>and</strong> substrate. [379] The<br />

low energy requirements of such manipulations make them<br />

inherently attractive for exploitation in future nanoelectronic<br />

devices [360t] but they have also inspired the design of structures<br />

specifically designed to exhibit mechanical switching of<br />

electrical properties induced by an STM tip. Molecules such<br />

as 118 (Scheme 71a) have been proposed in which two<br />

Scheme 71. a) Structure of “l<strong>and</strong>er” molecule 118 in which it is<br />

intended that manipulation of the angle between the phenyl rotator<br />

(red) <strong>and</strong> polyaromatic boards (blue) will control conductance through<br />

the molecule. [380] As for previous structures, interaction with the<br />

surface is through the orthogonally oriented di-tert-butyl “legs” (black).<br />

b) 1,4’’-Paratriphenyldimethylacetone (119) has been chemisorbed on<br />

a Si(100)-2 ” 1 surface through the oxygen atoms <strong>and</strong> the manipulation<br />

of the phenyl rings investigated by STM. [382] The arrows indicate the<br />

degrees of freedom which were probed with the tip.<br />

polyaromatic regions (blue) are connected by an aryl rotator<br />

(red). [380] In analogy to a series of rigid predecessors, [360t,381] the<br />

polyaromatic core of these molecular “l<strong>and</strong>ers” is intended to<br />

act as a single molecular wire,which is raised <strong>and</strong> insulated<br />

from the surface by the di-tert-butylphenyl legs (similar to<br />

116). If variation in the angle of the central rotator can<br />

significantly affect conductance through the molecule it<br />

would constitute a single-molecule nanomechanical electrical<br />

switch. A clear challenge here is achieving a system where<br />

thermal rotations of the central rotator do not erase the state<br />

of the switch while still allowing tip-induced rotation. A<br />

recent study of simple triphenyl 119 (Scheme 71b),chemisorbed<br />

through the oxygen atoms to a Si(100)-2 ” 1 surface<br />

has highlighted some of the problems facing such delicate<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

manipulations of intramolecular conformation with an STM<br />

tip. [382] However,reversible conformational switching of<br />

biphenyl molecules adsorbed on a Si(100) surface,has been<br />

successfully realized through electronic excitation using an<br />

STM tip. [383] Remarkably,two different dynamic processes<br />

could be initiated simply by positioning the tip at slightly<br />

different locations over the molecule.<br />

Investigations into the correlation <strong>and</strong> manipulation of<br />

mechanical modes within single molecules on surfaces has<br />

begun. These early studies highlight the exciting potential for<br />

successful working systems together with the difficulties<br />

involved in balancing the forces necessary for a molecule to<br />

overcome thermal motion while still allowing reversibly<br />

addressable internal modes. Pushing the 4-tert-butyl “h<strong>and</strong>les”<br />

(dark blue) with an STM tip was intended to rotate the<br />

triptycene “wheels” (red) <strong>and</strong> roll “molecular wheelbarrow”<br />

120 (Scheme 72) across a surface. [384,385] Simulations of early<br />

designs did not result in the intended motion on account of a<br />

Scheme 72. Structure of “molecular wheelbarrow” 120, designed to<br />

convert a directional translational stimulus applied by the STM tip at<br />

the “h<strong>and</strong>les” (dark blue) into directional rotation of the “wheels”<br />

(red). [384–386]<br />

flattening of the rear legs on the surface. [384a] If this effect was<br />

eliminated from the simulations,rotation of the wheels still<br />

did not occur,they simply glided across the surface. On the<br />

other h<strong>and</strong>,when second-generation system 120 was adsorbed<br />

on Cu(100),contact with the STM tip fragmented the<br />

molecule,thus suggesting strong interactions between the<br />

molecule <strong>and</strong> the surface even at room temperature. [386]<br />

Clearly a macroscopic wheelbarrow operates on the<br />

principle that friction between the wheel <strong>and</strong> the ground is<br />

greater than resistance to rotation around the axle. Friction is<br />

a function of the nature of the surfaces in contact <strong>and</strong> the<br />

weight of the barrow. At the molecular level,however,the<br />

concept of friction is less easily understood: [387] the resistance<br />

to sliding motion arises from individual short-range attractive<br />

<strong>and</strong> repulsive forces,as well as longer-range attractive van der<br />

Waals interactions,while the effect of gravity is negligible.<br />

Rolling translation on surfaces <strong>and</strong> the role of friction have<br />

136 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

been considered more extensively in relation to fullerenes<br />

<strong>and</strong> carbon nanotubes (see also Section 5) on account of their<br />

spherical <strong>and</strong> cylindrical shapes. [388] It has now been demonstrated<br />

experimentally that both carbon nanotubes [389] <strong>and</strong><br />

C 60 [390] can undergo tip-induced surface translations through a<br />

rolling mechanism in preference to sliding motions. This has<br />

recently been exploited by Tour <strong>and</strong> co-workers in a<br />

remarkable demonstration of what they term a singlemolecule<br />

“nanocar”, 121 (Figure 38a). [391] The molecule<br />

consists of an oligo(phenylethylene) “chassis” supporting<br />

four C 60-derived “wheels”. These molecules can be imaged on<br />

Au(111) by STM as four bright lobes that are stationary at<br />

room temperature. Increasing the substrate temperature<br />

results in 2D movement of the molecules,observed through<br />

sequential STM images (Figure 38c–g). The orientation of the<br />

molecules can be determined because of the difference in the<br />

length <strong>and</strong> width dimensions,<strong>and</strong> translational motion clearly<br />

occurs perpendicular to the axles,with pivoting around one<br />

wheel accounting for changes in direction. This orientation<br />

strongly suggests a thermally driven rolling mechanism,<strong>and</strong><br />

was confirmed by the behavior of two three-wheeled analogues,one<br />

of which (122,Figure 38b) was designed only to<br />

undergo pivoting,the other of which could not sustain any<br />

motions involving concerted rolling of all the wheels (not<br />

shown). Movement of 121 could also be achieved by<br />

manipulation with the STM tip,but again only motion<br />

perpendicular to the axles was facile. Intriguingly,this<br />

motion was generated with the tip in front of the molecule,<br />

pulling it along,in contrast to most other examples of tipinduced<br />

molecular movements. The authors propose that<br />

Figure 38. a) Chemical structure of four-wheeled “nanocar” 121. [391] b) Chemical structure of the three-wheeled analogue 122 with an axle<br />

arrangement designed to allow only pivoting motion through correlated rolling of the fullerene wheels. c)–g) STM images of 121 rolling on<br />

Au(111) at 2008C. The orientation of 121 can be determined by the wheel separation. Images (c)–(g) were selected from a series spanning<br />

10 min. h) Schematic representation of the motions of 121 <strong>and</strong> 122. For clarity the alkoxy substituents have been omitted. The STM images (c)–<br />

(g) <strong>and</strong> the schematic representation (h) are reprinted with permission from Ref. [391].<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

137


Reviews<br />

ultimately ensembles of such molecules could be propelled by<br />

external fields (see Section 5). [391]<br />

Cyclic molecules in which a position of unsaturation<br />

moves around equivalent positions through sigmatropic<br />

rearrangements (see Section 2.1.5) have been proposed as<br />

being able to roll across surfaces,powered by the breaking<br />

<strong>and</strong> formation of substrate–molecule bonds around the<br />

ring. [392] A related phenomenon has been observed experimentally<br />

by STM when a fluorinated C 60 derivative was<br />

deposited on Si(111)-(7 ” 7). The fullerenes were observed to<br />

move across the surface leaving a trail of fluorine atoms in<br />

their wake—presumably the result of a rolling motion<br />

powered by chemisorption of fluorine to the surface (somewhat<br />

reminiscent of the “chemical” Marangoni motion of<br />

droplets containing surface-active species,Section 6.1). [393]<br />

R<strong>and</strong>om surface motions of adsorbates can be directed<br />

along specific directions on low-symmetry surfaces. [394]<br />

Recently,however,STM was employed to observe the<br />

directional diffusion of a chemisorbed molecule (9,10dithioanthracene)<br />

on a high symmetry Cu(111) surface. [395]<br />

A size mismatch between the molecular dimensions <strong>and</strong><br />

surface periodicity results in sequential motion of the thiol<br />

“legs” so that the molecule s orientation is always maintained.<br />

Scanning probe microscopies also open up opportunities<br />

to design machines that operate <strong>and</strong> generate an output at the<br />

single-molecule level without being powered by direct<br />

manipulation by the probe tip. Electroactive organometallic<br />

rotor 123 (Figure 39) has been designed to function on a<br />

surface,attached through derivatization of the hydrotris(indazolyl)borate<br />

lig<strong>and</strong> (red) leaving the cyclopentadienyl<br />

lig<strong>and</strong> (blue) free to rotate. [385,396] The proposed mechanism<br />

for driven directional rotation envisages a single-molecule<br />

electronic junction through which selective oxidation of one<br />

ferrocene moiety occurs. Anodic repulsion of the resulting<br />

cation moves the charge towards the cathode where it is<br />

reduced,the rotation bringing the next ferrocene unit close to<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

the positive terminal for oxidation. It is not yet clear if such an<br />

experimental set-up is achievable or whether direct electron<br />

tunneling from the cathode to the anode would result instead<br />

or even whether thermal motion would scramble the rotor s<br />

position. Indeed,the low-energy barrier demonstrated for<br />

rotation around the ruthenium–cyclopentadienyl bond [397]<br />

requires that the electrostatic driving forces be significant at<br />

all times during the mechanism to avoid thermal scrambling.<br />

An alternative approach may be to bias rotation using a<br />

perpendicularly aligned electric field.<br />

Other techniques besides scanning probe microscopies<br />

allow the application of external fields (Section 5) for both<br />

imaging <strong>and</strong> manipulation with molecular-scale precision.<br />

Magnetic tweezers,glass microneedles,<strong>and</strong>,perhaps most<br />

prominently,optical tweezers have been used to investigate<br />

the mechanical behavior of biomolecules. [398] Many of the<br />

most significant advances with optical tweezers have been in<br />

the study of molecular motor proteins,the optical trap<br />

allowing monitoring <strong>and</strong> application of forces to motors <strong>and</strong><br />

their subunits at various stages in their operation. This<br />

technique has also been applied to protein folding <strong>and</strong><br />

unfolding,binding events,<strong>and</strong> DNA transcription. One of the<br />

most striking demonstrations of the capabilities of this tool is<br />

the tying of a knot in an actin str<strong>and</strong>,thereby determining<br />

significant mechanical properties of the filament. [399]<br />

Recently,a light pattern incident on an photoconductive<br />

surface was used to set up non-uniform electric fields which<br />

allow parallel trapping <strong>and</strong> manipulation of multiple particles<br />

over a large area—essentially a molecular-level example of<br />

dielectrophoresis. [400] Various methods for detecting single<br />

fluorophore molecules have been developed,thus leading to<br />

new insights in many biophysical investigations. [401] Together,<br />

all these tools for probing dynamic events,interactions,<strong>and</strong><br />

processes at the single-molecule level have had a profound<br />

effect on the underst<strong>and</strong>ing of many biological systems. The<br />

ability to examine single entities as opposed to ensemble<br />

Figure 39. a) Chemical structure of a potential molecular motor designed to operate at the single-molecule level. [385, 396] b) Proposed mechanism of<br />

operation: 1) a potential difference applied across a nanojunction results in oxidation of the ferrocene group nearest the anode (2); electrostatic<br />

repulsion results in rotation so as to bring the ferrocenium cation towards the cathode where it can be reduced (3); simultaneously a new<br />

ferrocene unit is brought close to the anode where it is oxidized (4), to generate a rotational isomer of (2). Fc = ferrocene; Fc + = ferrocenium.<br />

138 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

averages has allowed the study of phenomena in a manner not<br />

possible with a nonsynchronized population of molecules. We<br />

look forward to the increased application of these powerful<br />

techniques to synthetic molecular motors <strong>and</strong> machines over<br />

the next few years. On the other h<strong>and</strong>,as the design <strong>and</strong><br />

mechanistic complexity of synthetic devices grows,so more<br />

dem<strong>and</strong>ing questions will be asked of analytical techniques.<br />

Accordingly,the quest for new techniques for addressing <strong>and</strong><br />

analysis on the nanoscale continues apace. Very recently,for<br />

example,the creation of an optical “superlens” was reported<br />

which breaks the diffraction limit in optical microscopy <strong>and</strong><br />

allows imaging of objects significantly smaller than the<br />

wavelength of light used. [402] The resolution of the new lens<br />

is 60 nm,but this work points the way for progression towards<br />

truly nanoscale optical microscopy.<br />

8. From Laboratory to Technology: Towards Useful<br />

<strong>Molecular</strong> <strong>Machines</strong><br />

8.1. The Current Challenges: Constraining, Communicating,<br />

Correlating<br />

Although the development of reaction-driven,fielddriven,self-propelled,<strong>and</strong><br />

scanning-probe-controlled submolecular<br />

motion is rapidly gathering pace,it is the design <strong>and</strong><br />

synthesis of molecular structures—both classical covalent <strong>and</strong><br />

mechanically interlocked—which restrict the thermal movements<br />

of various submolecular components that has been<br />

central to the major advancements in molecular-level<br />

machines to date. These structural constraints have been<br />

successfully combined with external switching of molecular<br />

<strong>and</strong>/or electronic structure to create systems which exhibit a<br />

remarkable level of control over both the relative position of<br />

units as well as the frequency <strong>and</strong> even directionality of their<br />

motion. Whatever the application or the precise mode of<br />

action,however,it is clear any practical device requires that<br />

the molecular machine is able to interact with the outside<br />

world,either directly through macroscopic or nanoscopic<br />

property changes or through further interactions with other<br />

molecular-scale devices. The problem of how to “wire”<br />

molecular machines together <strong>and</strong> to the outside world also<br />

has implications for the physical construction of the devices<br />

which,in turn,puts further dem<strong>and</strong>s on the fidelity of the<br />

molecular-level mechanical processes over a range of conditions.<br />

In the following sections,we examine research that<br />

addresses these challenges by focusing on using molecularlevel<br />

motion to produce a functional property change or<br />

perform a physical task.<br />

Nature s motors <strong>and</strong> machines generally work at interfaces<br />

or on surfaces <strong>and</strong> the transfer of molecular-machine<br />

technology onto solid substrates is a key step in the development<br />

of many potential applications. Not only do solutionphase<br />

systems pose problems in terms of integration with<br />

current technologies,directly transmitting mechanical<br />

changes at the molecular level to the macroscopic world will<br />

generally require organization <strong>and</strong> cooperation of many<br />

molecular-level machines. There are already several examples<br />

of switchable <strong>and</strong> detectable supramolecular recognition<br />

Angew<strong>and</strong>te<br />

Chemie<br />

events (including threading <strong>and</strong> dethreading of pseudorotaxanes)<br />

which occur at surfaces. [403,404] The current state-of-theart<br />

in this regard is well represented by Stoddart,Zink,<strong>and</strong><br />

co-workers,who have harnessed the photochemically triggered<br />

dethreading of a surface-mounted pseudorotaxane to<br />

release molecules trapped in nanoscopic pores on the same<br />

surface. [405]<br />

The problem of achieving controlled movement within<br />

surface-mounted or solid-state interlocked molecules is more<br />

dem<strong>and</strong>ing,however,with both components constrained to<br />

the solid support or interface at all times. The first evidence<br />

for externally stimulated submolecular motions of a catenane<br />

in the solid state was observed when a vacuum-evaporated<br />

thin film of a benzylic amide [2]catenane was subjected to<br />

oscillating electric fields <strong>and</strong> produced an unexpected secondorder<br />

nonlinear optical effect. [406,407] In earlier studies,thin<br />

films of a metal-templated polyrotaxane were deposited on a<br />

carbon electrode by electropolymerization of pyrrole<br />

rings. [408] Each rotaxane monomer contained the three binding<br />

motifs present in [2]catenate [Cu(98)] (Scheme 60a,<br />

Section 4.6.1). Electrochemical reduction of the pentacoordinate<br />

Cu II species resulted in pirouetting of the macrocycle<br />

around the thread to give the Cu I co-conformer in its lower<br />

energy tetrahedral geometry. The motion was slow,however,<br />

taking 24 h to go to completion,<strong>and</strong> oxidation back to Cu II did<br />

not result in any detectable submolecular motion (the dpp-<br />

Cu II !terpy-Cu II transition is known to be slower in solution,<br />

see above). The same kind of motion was attempted in a<br />

catenate attached to a gold surface through S Au bonds.<br />

However,in this case no motion could be observed at all. [409]<br />

Electrochemically induced shuttling has been carried out<br />

on self-assembled monolayers (SAMs) of [2]rotaxane 124 on<br />

gold wire immersed in an electrolyte solution (Scheme 73). [410]<br />

The shuttle features TTF (green) <strong>and</strong> DNP (red) stations,<strong>and</strong><br />

voltammetry measurements consistent with reversible translation<br />

of the macrocycle were observed on oxidation <strong>and</strong><br />

subsequent reduction of the TTF unit. A related [2]rotaxane<br />

organized in Langmuir films undergoes shuttling induced by<br />

adding a chemical oxidant to the aqueous subphase. [411]<br />

Evidence for the submolecular motion was garnered from<br />

comparison of the p-A isotherms of the shuttle film with those<br />

of control compounds. X-ray photoelectron spectroscopy<br />

(XPS) of the initial <strong>and</strong> the oxidized films transferred onto<br />

solid substrates as Langmuir–Blodgett (LB) monolayers also<br />

confirmed a change in the position of the ring which was<br />

consistent with shuttling.<br />

A different redox-active molecular shuttle was successfully<br />

oriented perpendicular to the surface of a titanium<br />

dioxide nanoparticle through attachment to a tripodal<br />

phosphonate unit at one end of the thread. [412] The use of<br />

nanoparticles as a support allowed characterization of the<br />

dynamic processes in the surface-mounted species by both<br />

1 H NMR spectroscopy <strong>and</strong> cyclic voltammetry. Redox-triggered<br />

shuttling of the macrocycle was observed,although not<br />

with the expected switching characteristics. Paramagnetic<br />

suppression spectroscopy (PASSY),a novel NMR technique,<br />

suggests that the shuttling process,at least in solution,in these<br />

examples is complicated by various folded conformations of<br />

the thread. [413]<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

139


Reviews<br />

Scheme 73. An electrochemically switchable [2]rotaxane immobilized as a SAM on a gold wire. [410]<br />

Taking advantage of the distinct “rotator” <strong>and</strong> “stator”<br />

design of the second-generation overcrowded alkene rotary<br />

motors (Scheme 22,Section 2.2),Feringa <strong>and</strong> co-workers<br />

have anchored these molecules on gold nanoparticles through<br />

two thioalkyl substituents appended to the base unit. [414] The<br />

photochemical <strong>and</strong> thermal rotation was monitored by CD<br />

spectroscopy <strong>and</strong> also by differentiation of the two thioalkane<br />

tethers with an isotopic label,followed by cleavage of the<br />

motor molecules from the surface at various stages <strong>and</strong><br />

analysis in solution by NMR spectroscopy. [415]<br />

8.2. Reporting Controlled Motion in Solution<br />

Like other types of molecular switch, [151] a change in<br />

property induced by submolecular motion in a molecular<br />

machine could potentially be used in an information-storage<br />

or other switching device. The fundamental requirements for<br />

mechanical systems are the same as their nonmechanical<br />

counterparts <strong>and</strong> include: 1) a useful <strong>and</strong> detectable difference<br />

between two states; 2) fatigue resistance; 3) stability of<br />

each state under the operating conditions; 4) fast response<br />

times; <strong>and</strong>,particularly for memory applications,5) nondestructive<br />

read-out. [151] The first of these is rather open to<br />

interpretation. A change in the chemical shift of an NMR<br />

signal,for example,may be considered by a chemist to be<br />

both easily detectable <strong>and</strong> useful—<strong>and</strong> there is no reason why<br />

such outputs might not be technologically relevant in certain<br />

applications one day. However,most of the examples highlighted<br />

in this section are chosen in view of the technological<br />

advantages of particular output modes: optical,electronic,<br />

<strong>and</strong> mechanical. This indeed reflects the research direction of<br />

the field of molecular machines over the past five years,in<br />

which the interaction of the machine with the outside world<br />

has become an increasingly important design element.<br />

8.2.1. Conformational Switches in Solution<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

The stimuli-induced cessation of conformational motions<br />

(“molecular brakes”,discussed in Sections 2.1.2 <strong>and</strong> 2.1.3) can<br />

be employed for switching applications. Glass <strong>and</strong> Raker have<br />

developed sensors by using a “molecular pinwheel” architecture<br />

(for example, 125,Scheme 74) which exploits positive<br />

allosteric binding effects achieved in molecular brakes with<br />

multiple binding sites (for example,metal porphyrinate<br />

s<strong>and</strong>wich compound 22,Scheme 7,Section 2.1.3). [416] Each<br />

trityl unit has one blade functionalized with a fluorophore <strong>and</strong><br />

the other two with suitable binding sites. The binding of an<br />

analyte which can bridge between the two trityl moieties<br />

freezes the rotational motion of the trityl groups with respect<br />

to each other. The remaining binding sites are now preorganized,thus<br />

favoring a second,cooperative,binding of the<br />

substrate. The mode of binding holds the two fluorophores in<br />

proximity so that an increase in the ratio of the excimer:monomer<br />

fluorescence can be measured. Pinwheel receptor 125<br />

is therefore able to detect short dicarboxylates selectively<br />

over monocarboxylates in aqueous media. [416d] It is proposed<br />

that this strategy could provide a general means to create<br />

cooperative sensors through tailoring the binding sites <strong>and</strong><br />

distance between them to be specific for a particular analyte.<br />

A different approach to mechanically operated sensors is to<br />

append linear or macrocyclic polyethers to polythiophenes or<br />

oligothiophenes. In some cases binding can result in a<br />

significant perturbation to the conformation of the p-conjugated<br />

system,which results in changes to the molecular<br />

electronic properties. [417–419]<br />

The possibility of conformational changes being transmitted<br />

over relatively long distances through fused cyclic<br />

frameworks was first noted by Barton in his seminal work on<br />

the conformational analysis of steroids. [420] An impressive<br />

example was recently reported by Clayden et al.,who<br />

140 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 74. An example of a molecular pinwheel receptor, 125. [416d] Cooperative binding between pairs of convergent binding sites freezes internal<br />

rotation, thereby bringing the fluorophores into proximity so that the ratio of excimer:monomer fluorescence increases. The combination of<br />

cooperative binding <strong>and</strong> a mechanically transmitted binding-induced change in the fluorescence spectrum allows the sensing of short<br />

dicarboxylates at low concentrations in aqueous solution.<br />

employed the conformational interactions of aromatic amides<br />

in 126 to transmit stereochemical information from a chiral<br />

group to a prochiral reaction site over 25 Š <strong>and</strong> a minimum of<br />

23 covalent bonds (Scheme 75). [421]<br />

Such conformational communication also presents opportunities<br />

to relay a stimuli-induced conformational change—<br />

triggered by a stimulus at some receptor site—through some<br />

Scheme 75. Ultra-remote stereocontrol by conformational communication<br />

in the reaction of trisxanthene 126. [421] The chiral oxazolidine ring<br />

(red) imposes a right-h<strong>and</strong>ed (P) twist on the adjacent amide as a<br />

result of both electronic <strong>and</strong> steric interactions. As the amides in such<br />

xanthene derivatives strongly prefer all-anti conformations, the effect of<br />

the enantiopure unit is transmitted throughout the molecule as a<br />

strongly preferred conformation for all the amide groups. The result is<br />

that Grignard addition to the aldehyde (green) proceeds with > 95:5<br />

diastereofacialselectivity. Subsequent cleavage of the oxazolidine<br />

affords an almost enantiomerically pure (> 90 % ee) alcohol.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

structural framework so as to produce a detectable <strong>and</strong><br />

reversible signal at a remote reporter unit. Pyranose-based<br />

conformational switch 31,for example,(Scheme 11,Section<br />

2.1.4) has been adapted to bring together two fluorophores<br />

through a chelation-induced ring-flip,therefore altering<br />

the optical properties. [422] Similarly,the W-shaped transoid–transoid<br />

to U-shaped cisoid–cisoid switching of terpyridine<br />

derivatives (see Scheme 16,Section 2.1.4) has been<br />

utilized to create another cation-triggered optomechanical<br />

switch 127 (Scheme 76). [423] This same strategy has also been<br />

used to switch the distance between two appended porphyrins,thus<br />

controlling energy- <strong>and</strong> electron-transfer processes<br />

between them. [424]<br />

One challenge,however,is to transduce a signal over<br />

longer distances, <strong>and</strong> for that purpose, 2,3,6,7-tetrasubstituted<br />

perhydroanthracene 128 was designed <strong>and</strong> constructed. [425]<br />

Conformational switching is triggered by metal-ion binding to<br />

the bipyridyl receptor units (red). The resultant three-ring flip<br />

separates the pyrene reporter units (blue),thus causing a<br />

change in fluorescence (Scheme 77). The binding signal is<br />

relayed over a distance of about 2 nm,while an analogous<br />

system based on 2,3,6,7-tetrasubstituted cis-decalins operates<br />

over a shorter distance of 1.5 nm. [426] In a first step towards<br />

transmitting signals over longer distances,the coupling of two<br />

such cis-decalin switches has been reported: the conformational<br />

change in one unit precipitates flipping of the remote<br />

ring system. [427]<br />

The conformational bistability of resorcin[4]arene cavit<strong>and</strong>s<br />

(Scheme 12,Section 2.1.4) has been exploited to create<br />

a well-defined mechanical switch, 129,in which the distance<br />

between the extremities of the two “arms” is varied from<br />

approximately 0.7 nm to about 7 nm by a pH change<br />

(Scheme 78). [428] The conformational change is used to alter<br />

the distance between a donor–acceptor dye pair,thereby<br />

eliciting a major difference in the fluorescence resonance<br />

energy transfer (FRET) response.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

141


Reviews<br />

Scheme 76. A cation-triggered optomechanical switch. [423] When in the U-shape conformation, charge transfer from the pyrene fluorophores to the<br />

electron-poor complexed terpyridine (terpy) lig<strong>and</strong> effectively quenches fluorescence. Pulses of Zn II ions are provided by protonation <strong>and</strong><br />

deprotonation of the N,N-bis(2-aminoethyl)ethane-1,2-diamine (tren) lig<strong>and</strong>, so that, overall, the system is switched by alternate acid (CF 3SO 3H)<br />

<strong>and</strong> base (LiOH) stimuli.<br />

Scheme 77. Signal transduction by transmission of a conformational change from a receptor unit (red) to a reporter (blue). [425]<br />

8.2.2. Co-conformational Switches in Solution<br />

The principles of shuttling in metal-templated rotaxanes<br />

have been extended to create a dimeric system in which the<br />

submolecular motion results in lengthening <strong>and</strong> contraction<br />

of the molecule in a manner reminiscent of the operation of<br />

the actin–myosin complex,which is the basis of natural<br />

muscle. [191v,429] Each monomer (130,Figure 40) consists of a<br />

bidentate dpp site incorporated in a macrocycle (see 98,<br />

Section 4.6.1) connected to a thread which also contains a 2,9dimethylphenanthroline<br />

(dmp) lig<strong>and</strong> <strong>and</strong> a tridentate terpyridine<br />

(terpy) site. The Cu I ions used to template formation<br />

of the dimer coordinate to the dpp unit of one monomer <strong>and</strong><br />

the dmp moiety of the other to give the threaded structure<br />

[Cu 2(130) 2] 2+ (Figure 41). Unfortunately,electrochemical<br />

oxidation to the Cu II species did not trigger a change in coconformation;<br />

even this unfavorable geometry for divalent<br />

copper ions is too kinetically stable. However,demetalation<br />

(excess KCN,room temperature,3 h) to give the free lig<strong>and</strong><br />

system 130 2,followed by insertion of Zn II ions (Zn(NO 3) 2,<br />

room temperature,1 h) generated the contracted form<br />

[Zn 2(130) 2] 4+ ,a length change of approximately 85 Š to<br />

65 Š (a reduction of 24 %). The molecule could be returned<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

to its original length simply by treatment with excess<br />

[Cu(CH 3CN) 4]PF 6.<br />

Simple monomeric stimuli-responsive molecular shuttles<br />

offer a generic approach to mechanical molecular switches for<br />

distance-dependant properties through suitable functionalization<br />

of the macrocycle <strong>and</strong> one end of the thread<br />

(Figure 42). [430]<br />

Following the observation that the positioning of an<br />

intrinsically achiral benzylic amide macrocycle in relation to a<br />

chiral center can result in an induced circular dichroism<br />

(ICD) effect (Section 4.3.1),chiroptical molecular shuttle<br />

(E)/(Z)-131 was prepared (Figure 43). [431] Unlike chiroptical<br />

switches in which the presence of,or h<strong>and</strong>edness of,chirality<br />

is intrinsically altered, [150] (E)/(Z)-131 remains chiral <strong>and</strong><br />

nonracemic with the same h<strong>and</strong>edness throughout; it is the<br />

expression of chirality that is altered. In the (E)-131 form,the<br />

macrocycle is held over the fumaramide template <strong>and</strong> thus far<br />

from the chiral center of the peptidic station. Correspondingly<br />

the circular dichroism response is zero,as observed for the<br />

free thread. In the (Z)-131 maleamide isomer,however,the<br />

olefin hydrogen-bonding station is “switched off”,the macrocycle<br />

resides on the chiral peptide station,<strong>and</strong> a strong<br />

( 13 000 degcm 2 dmol 1 ) negative ICD response is observed.<br />

142 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 78. pH-Governed conformational switching to control a FRET response in resorcin[4]arene cavit<strong>and</strong> 129. [428] R = n-Hex.<br />

Figure 40. Monomer unit 130 for the construction of an artificial<br />

molecular muscle rotaxane dimer. [429]<br />

The E!Z isomerization is most efficiently carried out by<br />

irradiation at 350 nm in the presence of a benzophenone<br />

sensitizer (photostationary state 70:30 Z/E),while the Z!E<br />

transformation can be achieved almost quantitatively by<br />

irradiation at 400–670 nm in the presence of catalytic Br 2.<br />

Although a more modest difference in photostationary states<br />

is achieved by irradiation at 254 nm (photostationary state<br />

56:44 Z/E) <strong>and</strong> 312 nm (photostationary state 49:51 Z/E),a<br />

large net change (> 1500 deg cm 2 dmol 1 ) in the elliptical<br />

polarization response is still observed (Figure 43 b) <strong>and</strong> this is<br />

reproducible over several cycles without addition of any<br />

external chemical reagents. [431]<br />

Figure 41. Reversible switching between extended ([Cu 2(130) 2] 2+ ) <strong>and</strong><br />

contracted ([Zn 2(130) 2] 4+ ) forms in a chemically switched artificial<br />

molecular muscle. [429]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

A similar approach has been used to create a molecular<br />

shuttle switch for fluorescence,namely,(E)/(Z)-132<br />

(Figure 44). [430] This system also relies on the photoswitchable<br />

fumaramide/maleamide station,but attached to the intermediate-strength<br />

dipeptide station is an anthracene fluorophore<br />

while the macrocycle now contains pyridinium units,<br />

which are known to quench anthracene fluorescence by<br />

electron transfer. Strong fluorescence (l exc = 365 nm) is<br />

observed in both the free thread <strong>and</strong> (E)-132,while shuttling<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

143


Reviews<br />

Figure 42. Exploiting a well-defined, large amplitude positional change<br />

to trigger property changes. [430] a) A <strong>and</strong> B interact to produce a<br />

physical response (fluorescence quenching, specific dipole or magnetic<br />

moment, nonlinear optical properties, color, creation/concealment of a<br />

binding site or reactive/catalytic group, hydrophobic/hydrophilic<br />

region, etc.); b) moving A <strong>and</strong> B far apart mechanically switches off<br />

the interaction <strong>and</strong> the corresponding property effect.<br />

Figure 43. a) Chiroptical switching in the [2]rotaxane-based molecular<br />

shuttle (E)/(Z)-131. [431] b) Percentage of (E)-131 in the photostationary<br />

state (from 1 H NMR data) after alternating the irradiation between<br />

254 nm (half integers) <strong>and</strong> 312 nm (integers) for five complete cycles.<br />

The right-h<strong>and</strong> y-axis shows the CD absorption at 246 nm.<br />

of the macrocycle onto the glycylglycine station in (Z)-132<br />

almost completely quenches this emission. At the wavelength<br />

of maximum emission from (E)-132 (l max = 417 nm) there is a<br />

remarkable 200:1 difference in intensity between the two<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 44. A fluorescent molecular switch based on [2]rotaxane molecular<br />

shuttle (E)/(Z)-132. [430] a) Interconversion between fluorescent (E)-<br />

132 <strong>and</strong> nonfluorescent (Z)-132; b) images of cuvettes containing<br />

solutions of (Z)-132 <strong>and</strong> (E)-132, respectively (0.8 mm, CH 2Cl 2), illustrating<br />

the clearly visible difference in fluorescence intensity. The<br />

photograph was taken while illuminating the samples with UV light<br />

(365 nm).<br />

states which is strikingly visible to the naked eye (Figure<br />

44b). [432]<br />

A related strategy has been applied by Tian <strong>and</strong> coworkers<br />

to achieve fluorescence switching in a different type<br />

of [2]rotaxane system,(E)/(Z)-133 (Scheme 79). [433] In (E)-<br />

133·2 H,the a-cyclodextrin ring sits preferentially over the<br />

trans-stilbene unit <strong>and</strong> this co-conformation is apparently<br />

further stabilized by strong hydrogen-bonding interactions<br />

between hydroxy groups on the cyclodextrin 3-rim <strong>and</strong> the<br />

isophthaloyl stopper group. [434] As for previous a-CD-based<br />

rotaxanes (Scheme 33,Section 4.3.1),photoisomerization of<br />

the stilbene unit can only occur when thermal motion has<br />

moved the ring away from the binding site. However,the<br />

strength of the combined hydrophobic <strong>and</strong> hydrogen-bonding<br />

interactions (even in water) for (E)-133·2H means that the<br />

shuttle is effectively conformationally locked: irradiation at<br />

355 nm,which should isomerize the stilbene moiety,results in<br />

no change to the system. Formation of the disodium salt of the<br />

isophthaloyl group (giving (E)-133·2Na) breaks the hydrogen-bonding<br />

network <strong>and</strong> although NMR studies show that<br />

the cyclodextrin continues to sit over the stilbene station,<br />

enough thermal motion is now present to allow photoisome-<br />

144 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 79. Photoswitched shuttling in [2]rotaxane (E)/(Z)-133·2Na which results in fluorescence enhancement<br />

of the 4-aminonaphthalimide group. [433] The photoisomerization process is prevented when the free<br />

carboxylic acids of the isophthaloyl stopper are present ((E)-133·2H).<br />

rization to give (Z)-133·2Na (photostationary state 63:37<br />

Z/E). In the Z isomer the cyclodextrin ring is forced to reside<br />

over the biphenyl group. The translocation of the ring is<br />

accompanied by a 46 % increase in fluorescence intensity of<br />

the 4-aminonaphthalimide stopper (l max = 530 nm). This<br />

change is attributed to a restriction of the vibrational <strong>and</strong><br />

rotational movement of the methylene <strong>and</strong> biaryl linkages<br />

thus disfavoring nonradiative relaxation processes. The shuttling<br />

<strong>and</strong> fluorescence changes are reversible on reisomerization<br />

to (E)-133·2 Na at 280 nm.<br />

Using either a stilbene [50k] or an azobenzene [435] configurational<br />

switch,the same research group have extended this<br />

concept to [2]rotaxanes in which the fluorescence of either<br />

one of two different fluorophores can be enhanced depending<br />

on the position of the ring. [436] Incorporating two different<br />

configurational switches into such a rotaxane has allowed the<br />

creation of a molecular system capable of carrying out the<br />

function of a half-adder—a two-input,two-output device<br />

which combines AND <strong>and</strong> XOR logic functions <strong>and</strong> which<br />

can perform binary addition. [355] [2]Rotaxane 134 can be<br />

reversibly switched between four different configurational<br />

isomers through selective photochemical <strong>and</strong> thermal stimuli<br />

(Scheme 80). For operation as a logic gate,UV irradiation at<br />

380 nm <strong>and</strong> 313 nm can be considered as two input signals (I1<br />

<strong>and</strong> I2). In both (Z N=N,E C=C)-134 <strong>and</strong> (E N=N,Z C=C)-134,the<br />

proximity of the ring to one or other stopper increases its<br />

fluorescence by an appreciable amount,while in both (E,E)-<br />

134 <strong>and</strong> (Z,Z)-134 the cyclodextrin is not held close to either<br />

fluorescent stopper (because of rapid shuttling in the former<br />

case <strong>and</strong> trapping on the biphenyl unit in the latter).<br />

Therefore,if the change in fluorescence intensity relative to<br />

(E,E)-134 (DF) is taken as an output (O1),this can be<br />

interpreted as an XOR gate (see truth table in Scheme 80).<br />

All four states exhibit absorbance maxima at 270 nm,which is<br />

found to increase in intensity on successive E!Z isomer-<br />

ization of the double bonds,<br />

<strong>and</strong> at 350 nm,which<br />

decreases as E!Z for each<br />

configurational switch. The<br />

result is that the difference<br />

in the absorption at these two<br />

wavelengths,expressed as a<br />

fraction of the absorption at<br />

the isobestic point (301 nm),<br />

is relatively small for all but<br />

the (Z,Z)-134 isomer <strong>and</strong>,<br />

thus,this value can be considered<br />

the output of an AND<br />

gate (O2). As these two functions<br />

operate in parallel,are<br />

based on the same two inputs,<br />

<strong>and</strong> give two distinct outputs,<br />

this system represents monomolecular<br />

realization of a<br />

half-adder function operating<br />

purely through photonic stimuli.<br />

[252,355] It is also possible to<br />

construct a [3]rotaxane analogue<br />

of 134. [437] The resulting<br />

shuttle exhibits three stable states (E,E-, Z,E-,<strong>and</strong> E,Z-) each<br />

of which exhibits a different fluorescent response as a<br />

consequence of the position of the rings.<br />

The nature of interlocked architectures—in which certain<br />

functional groups are encapsulated by the binding cavity of a<br />

separate unit—suggests that their dynamic attributes may<br />

allow the development of novel applications in catalysis. As a<br />

first step towards creating a fully processive catalyst,Rowan<br />

<strong>and</strong> co-workers have created a macrocyclic unit incorporating<br />

a manganese porphyrin,which can catalyze the epoxidation of<br />

an olefin bound in the inner cavity when an external oxidant<br />

<strong>and</strong> a bulky lig<strong>and</strong> are also present. [438] Threading this unit<br />

onto a polybutadiene chain <strong>and</strong> stoppering to create a<br />

rotaxane architecture creates a mechanically linked catalyst–substrate<br />

system; when the external reagents are added,<br />

the macrocycle promotes epoxidation at multiple positions on<br />

the thread,although at present the movement <strong>and</strong> action of<br />

the macrocycle is presumably r<strong>and</strong>om <strong>and</strong> nondirectional<br />

between thread sites. [439] The mechanically linked nature of<br />

such catalyst–substrate complexes could lead to processive<br />

catalysts for the post-polymerization modification of polymers<br />

<strong>and</strong> provide insight into the operation of the processive<br />

enzymes which are so central to fundamental processes in the<br />

cell.<br />

8.3. Reporting Controlled Motion on Surfaces, in Solids, <strong>and</strong><br />

Other Condensed Phases<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Early successes at operating molecular machines on<br />

surfaces <strong>and</strong> at interfaces were discussed in Section 8.1.<br />

Incorporation of mechanical switches (Section 8.2) into such<br />

environments can lead to switchable solids <strong>and</strong> interfaces in<br />

which stimuli-induced molecular-level motions can be communicated<br />

to the macroscopic world as changes in the<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

145


Reviews<br />

Scheme 80. Operation <strong>and</strong> truth table for a molecular half-adder based on a fluorescent molecular shuttle. [355] Note, the E!Z photoisomerizations<br />

reached photostationary states, shown here as percentages of the major isomer over the reaction arrows. All the fluorescence <strong>and</strong><br />

absorption measurements were carried out on these photostationary state compositions <strong>and</strong> the differences in the DF <strong>and</strong> DA values were<br />

sufficient to discern “high” (1) <strong>and</strong> “low” (0) responses.<br />

properties of the material. At the very least,this requires the<br />

combined mechanical switching of an ensemble of molecules<br />

<strong>and</strong>,in the most sophisticated cases,coordinated coherent<br />

motions of a population of molecular machines is required.<br />

8.3.1. Solid-State <strong>Molecular</strong> Electronic Devices<br />

In a series of ground-breaking experiments that interface<br />

molecular-level machines with silicon-based electronics,<br />

molecular shuttles (Sections 4.2 <strong>and</strong> 4.3) have been employed<br />

as an active molecular component in solid-state molecular<br />

electronic devices. [440] In the first example of such devices to<br />

utilize interlocked molecules,a monolayer of redox-active<br />

degenerate two-station rotaxanes (135 4+ ,Scheme 81) was<br />

s<strong>and</strong>wiched between electrodes made from titanium <strong>and</strong><br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

aluminum oxide. [441] The resulting junctions exhibited a<br />

strongly nonlinear current–voltage response at small negative<br />

biases. Application of an oxidizing voltage however resulted<br />

in an irreversible reduction in conductance. Connection of<br />

these fuselike switches in linear arrays created wired-logic<br />

gates performing OR <strong>and</strong> AND functions that displayed more<br />

pronounced voltage responses compared to resistor-based<br />

circuits,<strong>and</strong> thus demonstrating the defect-tolerance of this<br />

architecture for generating logic functions.<br />

In a second generation of devices,bistable interlocked<br />

molecules were employed <strong>and</strong> reversible switching achieved.<br />

Structures containing the tetracationic cyclophane CBPQT 4+<br />

can be ordered in Langmuir films by using an amphiphilic<br />

counterion <strong>and</strong> the monolayers subsequently transferred onto<br />

solid substrates by using the LB technique. [442] The process<br />

146 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 81. Chemical structure of amphiphilic [2]rotaxane 135 4+<br />

employed in the first electronic switches based on an interlocked<br />

molecule. [441] Alignment of the molecules was achieved by exploiting<br />

the hydrophobic (black stoppers) <strong>and</strong> hydrophilic (light blue hydroxymethyl<br />

moiety) portions at the top <strong>and</strong> bottom of the molecule as<br />

drawn.<br />

was used to create a molecular switch tunnel junction (MSTJ)<br />

in which a monolayer of bistable [2]catenane 92 4+<br />

(Scheme 54,Section 4.6.1) was s<strong>and</strong>wiched between silicon<br />

<strong>and</strong> titanium/aluminum electrodes. [443] These devices exhibited<br />

a moderate (ca. 2 ” ) increase in conductance after<br />

application of an oxidizing potential,while recovery of the<br />

initial state occurred after a reducing voltage was applied. The<br />

“read” voltage was in the region 0.1 to 0.3 V <strong>and</strong> did not<br />

interfere with the switch configuration,while the devices<br />

withstood many switching cycles. A rational design process<br />

then led to devices made from a related [2]pseudorotaxane<br />

(136 4+ ,Scheme 82) in which hydrophobic <strong>and</strong> hydrophilic<br />

regions are now directly incorporated into the molecular<br />

structure to allow self-organization. [444] Although larger on/off<br />

current ratios <strong>and</strong> absolute currents were observed,the device<br />

characteristics were found to vary widely between both cycles<br />

<strong>and</strong> devices. On further investigation,these problems were<br />

attributed to the formation of molecular domains,which<br />

means that the device characteristics were no longer determined<br />

by single-molecule properties.<br />

Further refinements of this shuttle structure produced<br />

amphiphilic bistable [2]rotaxanes 137 4+ <strong>and</strong> 138 4+ . <strong>Molecular</strong><br />

switch tunnel junctions of these rotaxanes possessed stable<br />

switching voltages around 2 <strong>and</strong> + 2 V with reasonable on/<br />

off ratios <strong>and</strong> switch-closed currents. [445] These favorable<br />

characteristics allowed preparation of nanometer-scale devices<br />

which displayed properties similar to the original micrometer-sized<br />

analogues,thus suggesting a molecular-level<br />

mechanism for the MSTJ operation. Furthermore,these<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Scheme 82. Chemical structures of amphiphilic bistable [2]pseudorotaxane<br />

136 4+ <strong>and</strong> [2]rotaxanes 137 4+ <strong>and</strong> 138 4+ used as the active components<br />

MSTJs. [444,445] Hydrophobic (black) <strong>and</strong> hydrophilic (light blue) regions are<br />

incorporated directly into the structures to allow self-assembly in Langmuir<br />

films without requiring additives.<br />

devices could be successfully connected to form a 2D crossbar<br />

circuit architecture. The circuit could be used as a reliable 64bit<br />

r<strong>and</strong>om access memory (RAM). The more dem<strong>and</strong>ing<br />

task of creating a logic circuit was also demonstrated by hardwiring<br />

1D circuits. A genuine 2D MSTJ-based crossbar logic<br />

circuit,however,will require junctions with features yet to be<br />

attained by using molecular systems,such as diode character<br />

<strong>and</strong> gain.<br />

A limited number of non-interlocked <strong>and</strong> nonswitchable<br />

control molecules (for example,a simple alkyl chain carboxylic<br />

acid,the free CBPQT 4+ ring,<strong>and</strong> related degenerate<br />

catenanes) have been tested in similar settings to provide<br />

supporting evidence that a molecular-level electromechanical<br />

mechanism is responsible for the switching observed in the<br />

rotaxane-based devices. [442c,443–445] The studies suggest that<br />

some form of bistability,together with the presence of the<br />

macrocycle,are necessary for the switching properties <strong>and</strong> are<br />

therefore consistent with a mechanism whereby electronically<br />

stimulated shuttling alters the junction conductance. However,devices<br />

made from a single station [2]rotaxane still<br />

exhibit conductance switching,albeit with a significantly<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

147


Reviews<br />

lower on/off ratio than the two-station analogues 137 4+ <strong>and</strong><br />

138 4+ . [444] Crucially,all the solid-state devices exhibit marked<br />

temperature dependence,which showed that at least one<br />

thermally activated step is involved in their operation. This is<br />

in line with the observation of a metastable state for shuttling<br />

in solution at low temperature (Section 4.3.4), [228f] as well in<br />

similar interlocked systems aligned in SAMs, [410] Langmuir<br />

films,<strong>and</strong> Langmuir–Blodgett monolayers (Section 8.1), [411]<br />

<strong>and</strong> dispersed in a solid-state polymer electrolyte (Section<br />

8.3.2). [446] By comparison of these systems,together<br />

with the performance of a different rotaxane analogue in<br />

solution,in a polymer electrolyte,<strong>and</strong> in MSTJs, [228i] it has<br />

been demonstrated that the ground-state thermodynamics of<br />

these shuttles are controllable by manipulation of chemical<br />

structure <strong>and</strong> are qualitatively independent of the surrounding<br />

medium. However,the nature of the environment does<br />

strongly affect the kinetics for relaxation of the non-equilibrium<br />

state as one would expect,<strong>and</strong> the experimentally<br />

determined values from each environment quantitatively<br />

support a common molecular-level switching mechanism<br />

based on shuttling motion. [191ad,228f,i] The proposed mechanism<br />

for operation of the solid-state electronic devices is summarized<br />

in Figure 45.<br />

Identification of the high conductance state as that with<br />

the cyclophane encircling DNP (state (d),Figure 45) is<br />

supported by quantum mechanical computational studies in<br />

which the I–V response was simulated for a model pseudo-<br />

Figure 45. Proposed mechanism for the operation of rotaxane-based MSTJs. [191x,y,ad,228f] The coloring of<br />

functional units corresponds to that used for structural diagrams in Scheme 82. a) In the ground state, the<br />

tetracationic cyclophane (dark blue) mainly encircles the TTF station (green) <strong>and</strong> the junction exhibits low<br />

conductance (the precise ratio of co-conformers in this state is dependent on the exact chemical structure<br />

of the [2]rotaxane <strong>and</strong> in some cases can also be temperature dependent). b) Application of a positive bias<br />

results in one- or two-electron oxidation of the TTF units (green!pink) <strong>and</strong> the resulting electrostatic<br />

repulsion causes shuttling of the ring onto the DNP station (red, c). d) Returning the bias to near 0 V<br />

reveals a high conductance state in which the TTF units have been returned to neutral, but translocation of<br />

the cyclophane has not yet occurred, because of a significant activation barrier. Thermally activated decay<br />

of this metastable state ((d)!(a)) may occur slowly, over time (dependant on temperature) or can be<br />

triggered by application of a negative voltage (e) which temporarily reduces the cyclophane to its diradiacal<br />

dication form (dark blue!orange), thus allowing facile recovery of the thermodynamically favored coconformation<br />

(f). A similar mechanism is applicable in the operation of devices based on analogous<br />

[2]catenanes.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

rotaxane in each of the two co-conformations (that is,<br />

corresponding to (a) <strong>and</strong> (d) in Figure 45). [447] The picture<br />

presented in Figure 45 is not entirely accurate,however,<br />

particularly with regard to the conformation of the rotaxane<br />

molecules. Analysis of the structures of Langmuir <strong>and</strong> LB<br />

films composed of various amphiphilic [2]rotaxanes indicates<br />

that unless high surface pressures or larger hydrophilic<br />

stoppers are employed,a very acute angle to the interface is<br />

adopted,probably as a result of the hydrophilicity of the<br />

tetracationic cyclophane. [448] These experimental results were<br />

recently reproduced by a molecular dynamics study of the<br />

Langmuir films,in which similar tilted arrangements were<br />

observed for both co-conformations. [449] The overall conformation<br />

of the thread portion is folded <strong>and</strong> dependent on the<br />

position of the macrocycle. Folded conformations resulting<br />

from “alongside” interactions between the CBPQT 4+ moiety<br />

<strong>and</strong> the vacant p-donor station are often observed in<br />

solution, [191ad, 217j,225e,228c,d,h,238] so it seems likely that such<br />

arrangements are also present at surfaces <strong>and</strong> in the solid<br />

state. The importance of conformation on electronic properties<br />

was highlighted in a computational study in which density<br />

functional theory was used to calculate the electronic<br />

structure of various functional components of the rotaxanes<br />

<strong>and</strong> from which a description of the rotaxane frontier<br />

molecular orbitals could be derived. [450] It was found that<br />

the ring-on-DNP co-conformation was only the most conductive<br />

if the cyclophane participates directly in electron<br />

tunneling by providing low-lying<br />

LUMO orbitals. Such a situation is<br />

most plausible for a folded conformation<br />

with direct interactions between<br />

the ring <strong>and</strong> the unoccupied station.<br />

The molecular dynamics simulation of<br />

Langmuir films, [449] as well as a similar<br />

treatment of thiol-anchored SAMs on<br />

Au(111), [451] suggest that such folded<br />

conformations form thermodynamically<br />

favored superstructures,while<br />

the added intra- <strong>and</strong> intermolecular<br />

interactions do not alter the relative<br />

stabilities of the two co-conformations.<br />

Such arrangements actually place the<br />

rotaxane functional units in relative<br />

orientations comparable to the analogous<br />

[2]catenane which originally displayed<br />

very similar device character-<br />

istics. Density functional calculations<br />

for optimized monolayers of [2]catenane<br />

92 4+ s<strong>and</strong>wiched between two<br />

gold electrodes demonstrated again<br />

that the ring-on-DNP co-conformation<br />

exhibits the higher conductance at<br />

small applied read bias,with the conduction<br />

involving HOMOs from TTF<br />

as well as DNP <strong>and</strong> LUMOs from<br />

CBPQT 4+ . [452]<br />

Clearly the most convincing verification<br />

of an electromechanical mechanism<br />

for conductance switching<br />

148 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

would be direct real-time observation of the shuttles in an<br />

operating solid-state device. Probing a single molecular<br />

monolayer confined between two electrodes is a challenge<br />

which cannot yet be routinely met by modern-day analytical<br />

methods,<strong>and</strong> preliminary attempts at using grazing-angle<br />

refection-absorption infrared spectroscopy (RAIRS) to study<br />

mechanical movements in such scenarios have proven unsuccessful.<br />

[453] Following on from indirectly observed shuttling in<br />

Langmuir monolayers [411] (see Section 8.1),the first direct<br />

evidence for shuttling in such a condensed phase was recently<br />

obtained by using X-ray reflectometry. [454] These results have<br />

also confirmed the considerably tilted <strong>and</strong> folded rotaxane<br />

conformations which were previously inferred (see above).<br />

Although these devices show potential as possible components<br />

for genuine single-molecule electromechanical<br />

switches,a significant barrier to achieving this goal is the<br />

nature of the physical connections used to wire the device. At<br />

very small dimensions,it is expected that metal wires will<br />

exhibit the best conductance characteristics,but unfortunately<br />

a range of molecular shuttle devices with two metal<br />

electrodes failed to demonstrate switching properties that<br />

were dependent on the nature of the organic monolayer. [455]<br />

Indeed,the polarity of molecule–metal interfaces <strong>and</strong> the<br />

potential for the formation of metal filaments within molecular<br />

monolayers are emerging as recurring issues in metalcontacted<br />

organic molecular electronics. [456] However,catenane-based<br />

MSTJs which utilize single-walled carbon nanotubes<br />

in place of the silicon electrode have been successfully<br />

created <strong>and</strong> this may provide a valuable alternative in the<br />

creation of real-world devices. [457] Whether the philosophy of<br />

using thermally activated nuclear motions which occur on<br />

timescales many orders of magnitude slower than electron<br />

motions is the right one for a practical molecular computer or<br />

not is not necessarily the most valuable aspect of this work.<br />

The interfacing of molecular machines with electronics opens<br />

up the possibility of many types of practical hybrid devices<br />

<strong>and</strong> will doubtless prove seminal in getting molecular motors<br />

<strong>and</strong> machines out of solutions in laboratories <strong>and</strong> into<br />

technological devices.<br />

8.3.2. Using <strong>Mechanical</strong> Switches to Affect the Optical Properties<br />

of Materials<br />

The configurational changes that occur in photochromic<br />

dopants have been used for a number of years to initiate<br />

changes in a number of liquid-crystal properties. [458] The<br />

trans!cis photoisomerization of stilbene or azobenzene<br />

dopants,for example,can cause reversible disruption of the<br />

nematic mesophase,effectively “turning off” liquid crystallinity.<br />

However,a more subtle form of mechanical molecularlevel<br />

control is achievable in systems where the mesophase<br />

structure is maintained throughout. It is possible to orient<br />

photochromic dopants in a liquid-crystalline phase through<br />

axis-selective photoconversion by applying plane-polarized<br />

light to stimulate repeated photoisomerization. This directional<br />

orientation triggers alignment of the whole liquidcrystal<br />

phase. A second vector of polarization can then be<br />

used to produce a differently aligned state. This type of<br />

process has recently been modeled as a Brownian ratchet—<br />

Angew<strong>and</strong>te<br />

Chemie<br />

each of the interconverting forms of the dopant molecule<br />

move over a different potential energy surface as a result of<br />

different interactions with the liquid-crystalline phase,while<br />

transitions between them are fueled by the optical stimulus.<br />

[459] Such systems clearly exhibit remarkable amplification<br />

of a relatively small photoinduced configurational change<br />

which causes a global orientation change in the bulk. It is even<br />

possible to create surface-controlled systems where a monolayer<br />

of photochromic species (a so-called “comm<strong>and</strong> surface”)<br />

controls the alignment of a whole liquid-crystal layer in<br />

contact with it. Potential applications of this photoinduced<br />

mechanical molecular-level motion include optical recording<br />

media <strong>and</strong> light-addressable displays. [458]<br />

Chiral dopants can induce the formation of chiral nematic<br />

(cholesteric) phases. If the dopants contain photoswitchable<br />

units,variations in the expression of chirality can be achieved.<br />

Recently a dopant of fixed chirality bound to an azobenzene<br />

photoswitch was used to reverse the screw sense of the<br />

cholestric phase. [460] Some chiroptical switches,such as the<br />

overcrowded alkenes discussed in Section 2.2,have a change<br />

in helicity associated with their photochromic switching <strong>and</strong><br />

so are well-suited to switching the cholesteric liquid-crystal<br />

chirality. [461] Unidirectional molecular motor 46 (Figure 18,<br />

Section 2.2) has also been shown to play such a role. [462] Of the<br />

three isomers which can be isolated at room temperature,<br />

each exerts a different helical twisting power on the liquidcrystalline<br />

matrix,with opposite dopant helical h<strong>and</strong>edness<br />

inducing contrary screw senses in the liquid crystal. Photochemical<br />

<strong>and</strong> thermal unidirectional rotation of the motor<br />

was shown to occur within the liquid-crystal matrix,albeit<br />

with reduced efficiency <strong>and</strong> rates compared to solution<br />

studies. Liquid-crystal films doped with (P,P)-trans-46 gave<br />

a cholesteric phase with right-h<strong>and</strong>ed helicity <strong>and</strong> displayed a<br />

violet color. Photoirradiation (l > 280 nm) at room temperature<br />

leads to a decrease in the proportion of this isomer <strong>and</strong><br />

concomitant increases in the amount of (P,P)-cis-46 (which<br />

has a much weaker right-h<strong>and</strong>ed helical-twisting power) <strong>and</strong><br />

(M,M)-trans-46 (which exerts a left-h<strong>and</strong>ed helical twist). The<br />

combined effect of the change in dopant isomer composition<br />

is to progressively change the color of reflected light from<br />

violet through to red (Figure 46). A blue shift of the reflected<br />

wavelength could also be achieved on heating the film to<br />

608C which converts any (M,M)-trans-46 present back into<br />

(P,P)-trans-46. [462] As a consequence of the induced helicity in<br />

the liquid-crystal phase,it seems likely that the reorganizational<br />

motions triggered by isomerization of the dopant<br />

molecule are directional in nature. The color change,<br />

however,is not a result of the unidirectional trajectory of<br />

the motor; rather it is an expression of the system s state at<br />

Figure 46. Color changes in a liquid-crystal film doped with 46<br />

(6.16 wt%). [462] Starting from 100% (P,P)-trans-46, the sample is<br />

irradiated with light of l > 280 nm at room temperature. The pictures<br />

of the film shown from left to right were taken at 0, 10, 20, 30, 40, <strong>and</strong><br />

80 s. Reprinted with permission from Ref. [462].<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

149


Reviews<br />

any particular moment in time,a function of the<br />

different populations of a three- (or four-) state<br />

switch (see also Section 8.3.3). In other words,in this<br />

situation the motor molecule is simply acting as a<br />

switch <strong>and</strong> the directionality of the rotation of its<br />

components probably plays no role in changing the<br />

liquid-crystal color.<br />

The use of controlled molecular motion in<br />

polymer films to generate patterns visible to the<br />

naked eye (Figure 47) has been demonstrated with a<br />

Figure 47. Images obtained by casting films of polymer 139 on quartz<br />

slides, covering them with an aluminum mask, <strong>and</strong> exposing the<br />

unmasked area to dimethyl sulfoxide vapors for 5 min. [463] The photographs<br />

were taken while illuminating the slides with UV light (254–<br />

350 nm). The symbols of Sony Playstation are illustrative of the types<br />

of patterns that can be created.<br />

molecular shuttle based fluorescent switch derivatized<br />

with poly(methyl methacrylate) (139,<br />

Scheme 83). [463] A polymer film INHIBIT logic gate<br />

based on a combination of stimuli-controlled submolecular<br />

positioning of the ring <strong>and</strong> protonation<br />

was also demonstrated (140/140·2 H + ,Scheme 83,<br />

Figure 48). [463]<br />

Scheme 83. Polymeric environment-switchable molecular shuttles 139<br />

<strong>and</strong> 140/140·2H + . [463]<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 48. A molecular-shuttle Boolean logic gate that functions in a polymer<br />

film. [463] a) The aluminum grid used in the experiment. The coin shown for scale<br />

is a UK 5p piece. b) The pattern generated when films of 140 were exposed to<br />

trifluoroacetic acid vapor for 5 min through the aluminum grid mask. c) The<br />

criss-cross pattern obtained by rotating the aluminum grid 908 <strong>and</strong> exposing the<br />

film shown in (b) to DMSO vapor for a further 5 min. Only regions exposed to<br />

trifluoroacetic acid but not to DMSO are quenched. The truth table for an<br />

INHIBIT logic gate is shown in the inset. The photographs of the slides were<br />

taken in the dark while illuminating with UV light (254–350 nm).<br />

Many of the switchable rotaxane <strong>and</strong> catenane systems<br />

which employ charge-transfer interactions between the components<br />

intrinsically involve changes in absorption profiles—<br />

<strong>and</strong> therefore color—concomitant with shuttling. Incorporation<br />

of a rotaxane similar to 124 4+ (Scheme 73) into a solidstate<br />

polymer electrolyte gives an electrochromic device in<br />

which a color change occurs on application of an oxidizing<br />

potential. [446] The observed color change is consistent with<br />

that seen on shuttling in solution for this family of catenanes<br />

<strong>and</strong> rotaxanes (for example,[2]catenane 92 4+ ,Scheme 54,<br />

Section 4.6.1),while on subsequent reduction,the reverse<br />

color change occurs slowly,<strong>and</strong> is indicative of the slow<br />

relaxation of the non-equilibrium system in the solid state.<br />

Both these observations serve as strong evidence for shuttling<br />

in the condensed phase (as discussed in Section 8.3.1).<br />

Stoddart,Goddard,<strong>and</strong> co-workers have recently proposed<br />

an extension of this concept,based on [2]catenane 92 4+ ,but<br />

incorporating a third,intermediate <strong>and</strong> electroactive,station<br />

to create a three-state switchable catenane in which each of<br />

the states should be characterized by a different color arising<br />

from the intercomponent charge-transfer interactions. [464]<br />

8.3.3. Using <strong>Mechanical</strong> Switches To Affect the <strong>Mechanical</strong><br />

Properties of Materials<br />

Materials which can transduce an external stimulus into a<br />

mechanical response are at present probably best exemplified<br />

by piezoelectrics,but the development of molecular materials<br />

150 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

which exhibit similar—or potentially even more sophisticated—properties<br />

is an area of intense interest. [465] It is well<br />

established that conformational changes in polymers can<br />

result in mechanical changes at the macroscopic level. Such<br />

effects have been extensively studied for over half a century in<br />

stimuli-responsive hydrogels. These water-swollen 3D-crosslinked<br />

networks of neutral or ionic (also known as polyelectrolytes<br />

in this case) amphiphilic polymers can undergo<br />

reversible phase transitions in response to a number of<br />

stimuli,including pH,temperature,solvent composition,<br />

electric fields,<strong>and</strong> light. The phase changes result from<br />

altering the balance between hydrophilic <strong>and</strong> hydrophobic<br />

forces <strong>and</strong> can be accompanied by volume changes of up to<br />

two orders of magnitude. [466] These were perhaps the first<br />

synthetic chemomechanical systems to be investigated <strong>and</strong><br />

their significance was recognized early on by Katchalsky <strong>and</strong><br />

co-workers,who developed a number of remarkable macroscopic<br />

devices for directly <strong>and</strong> continuously transducing<br />

chemical potential into mechanical work. [467,468] Recent developments<br />

include using photons to induce phase transitions, [469]<br />

the use of oscillating chemical reactions to control the<br />

environment for pH-responsive gels to produce autonomously<br />

fluctuating chemomechanical systems, [470] <strong>and</strong> “core–<br />

shell” architectures to give materials with multistep volume<br />

changes. [471] Synthesis <strong>and</strong> optimization of such systems in the<br />

future may also be expedited by a supramolecular approach<br />

to their construction. [472] A wide range of technologically<br />

relevant devices have been proposed <strong>and</strong> are under development<br />

using such materials,with functions including the<br />

binding <strong>and</strong> release of guests, [473] flow control valves for<br />

microfluidic devices, [474] a macroscopic walking device, [475]<br />

control of enzyme activity, [476] as well as a number of<br />

biomedical applications. [477] Incorporation of molecular recognition<br />

units into the gel network can lead to volume<br />

changes—<strong>and</strong>,in turn,other effects—upon binding of a<br />

specific guest molecule or ion. [478] This approach has not only<br />

been successful for relatively simple host–guest systems,but<br />

with redox-active enzymes,such as glucose oxidase. Binding<br />

to,<strong>and</strong> subsequent oxidation of,the substrate produces the<br />

charged form of the enzyme cofactor which initiates the gel<br />

phase transition. [478c,d]<br />

Another emerging area of molecular-level mechanical<br />

motion in which hydrogels <strong>and</strong> other responsive polymeric<br />

materials have been applied,is the development of shapememory<br />

materials,in particular ones which are biologically<br />

compatible. [479] While most systems use a temperature change<br />

to induce recovery of the “memorized” shape,light-induced<br />

shape-memory effects have recently been reported in which a<br />

photoinitiated radical reaction has been used to rearrange<br />

covalent cross-links [480] or the reversible formation of crosslinks<br />

through a [2+2] cycloaddition which can be induced <strong>and</strong><br />

reversed by irradiation with light of different wavelengths. [481]<br />

Stimuli-induced mechanical changes can also be induced<br />

in conducting polymers. Electrochemical oxidation <strong>and</strong><br />

reduction of polypyrrole,polythiophene,or polyaniline films<br />

induces counterion transport into or out of the polymer<br />

matrix,thus resulting in volume changes. [482] Such effects were<br />

first demonstrated for macroscopic devices immersed in<br />

aqueous electrolytes,but the long reaction times <strong>and</strong> the<br />

Angew<strong>and</strong>te<br />

Chemie<br />

small forces generated are practical limitations of such<br />

systems. More recently,solid-state, [483] gel, [484] <strong>and</strong> ionicliquid<br />

[485] electrolytes have been employed,while the development<br />

of microfabricated devices has been successful in<br />

realizing quite complex <strong>and</strong> potentially technologically relevant<br />

electrochemomechanical actuators. [482c] The combination<br />

of the conducting polymer technology with a hydrogel<br />

material has been used to overcome some of the problems<br />

associated with each individual approach to demonstrate a<br />

novel drug-delivery system, [486] while similar processes are<br />

now being investigated in conductive supramolecular crystalline<br />

materials. [487] A related but novel mechanism for electromechanical<br />

actuation,which avoids many of the drawbacks<br />

associated with redox-induced dopant diffusion,has been<br />

realized in sheets of entangled nanotube bundles [488] <strong>and</strong> in<br />

porous films of gold nanoparticles. [489] The process (termed<br />

double-layer charging) does not involve any redox reaction,<br />

but rather the injection of charge into these high surface area<br />

materials attracts electrolyte ions to the surfaces,thus<br />

resulting in surface-stress-induced deformation.<br />

In the preceding examples a macroscopic mechanical<br />

change is produced by the combined effect of a number of<br />

nonspecific conformational changes in the polymeric network.<br />

Even if in some cases this is the result of a specific<br />

recognition event,the mechanical change is not an intrinsic<br />

feature of the molecular components. There are a number of<br />

systems,however,in which particular submolecular conformational,co-conformational,<strong>and</strong>/or<br />

configurational switches<br />

can be used to directly trigger macroscopic motion. These<br />

approaches circumvent the problems of diffusion-limited<br />

rates <strong>and</strong> uniformity of response throughout the material.<br />

The incorporation of configurational switches,especially<br />

azobenzenes,to reversibly control secondary <strong>and</strong> tertiary<br />

structure in biopolymers <strong>and</strong> synthetic oligomers (Section<br />

2.2) or to control long-range order in liquid-crystalline<br />

phases (Section 8.3.2) has already been discussed. A similar<br />

approach can be used for synthetic polymers,where the effect<br />

is often a macroscopically detectable response.<br />

Several types of azobenzene-containing polymer films<br />

undergo contraction–extension cycles on photoisomerization,<br />

[490] while other physical properties such as viscosity <strong>and</strong><br />

solubility can be similarly controlled. [491] Length changes in an<br />

azobenzene-containing polymer have recently been observed<br />

at the single-molecule level using AFM. [492] One end of a<br />

polyazopeptide (a polymer of azobenzene units linked by<br />

dipeptide spacers) was isolated on a glass slide while the other<br />

was attached to a gold-coated AFM tip through a thioether–<br />

gold bond (Scheme 84). At zero applied force,the polymer<br />

showed repeated shortening <strong>and</strong> lengthening on photoisomerization<br />

of the azobenzene units to cis <strong>and</strong> trans,respectively.<br />

Furthermore,this experimental arrangement was used to<br />

demonstrate the transduction of light energy into mechanical<br />

work. A “load” was applied to one of the polyazopeptides—in<br />

its fully trans form—by increasing the force applied by the<br />

AFM tip from 80 pN to 200 pN,thus resulting in a stretching<br />

of the polymer. Photoisomerization to increase the percentage<br />

of cis linkages in the polymer str<strong>and</strong> resulted in<br />

contraction,even against the externally applied force (“raising<br />

the load”). Removal of the “load” (reducing the applied<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

151


Reviews<br />

Scheme 84. Experimental set-up for the observation of single-molecule extension <strong>and</strong><br />

contraction using AFM (the azobenzene unit is shown as the extended Z isomer). [492]<br />

force) followed by reisomerization to the trans form,restored<br />

the polymer to its original length,ready to lift another load.<br />

The work done in this way by the single-molecule optomechanical<br />

device was approximately 5 ” 10 20 J.<br />

Analogous to controlling the chirality of liquid-crystal<br />

phases,<strong>and</strong> following the observation of environment-sensitive<br />

screw-sense inversion in certain biopolymers,there is<br />

considerable interest in remotely controlling the h<strong>and</strong>edness<br />

of synthetic helical polymer systems. [493] While this has been<br />

achieved in a number of nonspecific ways (for example,by<br />

changing the solvent or temperature) molecular switches can<br />

also be employed. Chiral azobenzene side chains permit<br />

reversible switching between P <strong>and</strong> M helices by controlling<br />

the proximity of the chiral center to the polymer backbone.<br />

[494] Circularly polarized light has been used to generate<br />

an enantiomeric excess in an axially chiral alkene pendant<br />

group. [495] As observed for the liquid-crystal examples,even a<br />

very small enantiomeric excess is amplified into a bulk chiral<br />

rearrangement of the polymer. Thin films of liquid-crystalline<br />

or amorphous azobenzene side-chain polymers can also be<br />

reoriented on the nanoscale level with linearly polarized<br />

light. [496] The photoinduced anisotropy confers dichroism <strong>and</strong><br />

birefringence on the material,thus suggesting potential for<br />

application in holographic data storage. [497] Intriguingly,such<br />

optically induced birefringence gratings are created in a<br />

process which can also involve mass transport to give surface<br />

relief features detectable by AFM or even the naked<br />

eye. [498,499] Indeed,the properties of liquid crystals <strong>and</strong><br />

polymers can be combined in liquid–crystal elastomers—<br />

polymers constructed from mesogenic monomers which can<br />

become oriented in an ordered manner. The nematic to<br />

isotropic phase transition can result in a macroscopic<br />

response. [500] If an azobenzene configurational switch is also<br />

incorporated,switching between nematic <strong>and</strong> isotropic<br />

phases,or reorientation of the chromophores,can be achieved<br />

on irradiation with linearly polarized light,<strong>and</strong> quite specific<br />

changes in the elastomer film can be induced. [501] This has<br />

even been demonstrated in a system where the azobenzene<br />

photoswitch is not covalently attached to the main polymer<br />

backbone <strong>and</strong>,in this case,the light-triggered deformation<br />

can also propel the elastomer across a water surface. [501g,502]<br />

On doping a nonpolymeric liquid-crystal film with lightdriven<br />

unidirectional rotor 141 (Figure 49 a,a member of the<br />

“third generation” series discussed in Section 2.2),the helical<br />

organization induced by the dopant results in a fingerprintlike<br />

structure to the surface of the film. [503] Both thermal<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

relaxation steps can occur at room temperature<br />

with this motor molecule. Irradiation of the film<br />

with light alters the distribution of the isomers<br />

present <strong>and</strong>,as they have different helical twisting<br />

powers,the organization of the liquid crystal is<br />

changed. This results in a rotational reorganizaton<br />

of the surface structure,which can be followed by a<br />

glass rod sitting on the film (Figure 49b). After<br />

irradiation of the sample for ten minutes,however,<br />

the rod s rotary motion ceases,which is indicative<br />

of the photostationary state being reached.<br />

Although there continues to be a directional flux<br />

Figure 49. a) Chemical structure of light-driven unidirectional molecular<br />

rotor 141. [503] b) Rotation of a glass rod (5 ” 28 mm) on a liquidcrystal<br />

film doped with 141 (1 wt%). Pictures (from left to right) were<br />

taken at 0, 15, 30, <strong>and</strong> 45 s <strong>and</strong> show clockwise rotations of 0, 28, 141,<br />

<strong>and</strong> 2268, respectively. Scale bars: 50 mm. The pictures are reprinted<br />

with permission from Ref. [503].<br />

of motor molecules between the different isomers,the net<br />

distribution of the isomers is no longer changing (see<br />

Section 2.2) <strong>and</strong> so the reorganization of the liquid-crystal<br />

matrix ceases. Removing the light stimulus allows the<br />

population of the “unstable” isomer to decay,returning the<br />

system to its starting state,accompanied by rotation of the rod<br />

in the opposite direction. Similar to the color switching<br />

discussed in Section 8.3.2 (Figure 46),this effect is not a<br />

product of the directional trajectory of a motor. The liquidcrystal<br />

organization reflects the distribution of dopant<br />

isomers at any particular moment in time,while the directionality<br />

of the rod s movement is due to the original chirality<br />

of the liquid-crystalline phase <strong>and</strong> dopant.<br />

In terms of conducting polymers,it is possible that an<br />

intrinsic molecular-level change of length is a contributing<br />

factor to the mechanism of operation of polyaniline-based<br />

electrochemomechanical actuators. [504] However,new systems<br />

are being developed in which a designed molecular-level<br />

conformational change is responsible for inducing mechanical<br />

motion. These include a thiophene-fused [8]annulene which<br />

switches between puckered <strong>and</strong> planar forms depending on<br />

the oxidation state. [505] The conformational flexibility of<br />

calix[4]arenes has led them to be incorporated as hinge<br />

elements into conducting polymers based on oligothiophenes.<br />

[506] It has been proposed that electrochemical switching<br />

of p–p interactions between adjacent oligothiophene units<br />

152 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

should lead to microscopic <strong>and</strong> macroscopic mechanical<br />

motions,although there is some debate over the precise<br />

mechanism. [507]<br />

Although the metal-templated rotaxane dimer illustrated<br />

in Figure 41 (Section 8.2.2) exhibits reversible switching of its<br />

molecular length,cooperative coherent motion in a population<br />

of 130 2 to generate a macroscopic response has yet to be<br />

achieved. The first demonstration of macroscopic mechanical<br />

motion triggered by shuttling in a rotaxane was recently<br />

demonstrated using electroactive [3]rotaxane 142 8+ . [508] Oxidation<br />

of the TTF stations (green!pink) results in shuttling<br />

of the cyclophanes onto the hydroxynaphthalene units (red),<br />

significantly shortening the inter-ring separation (Scheme 85).<br />

A self-assembled monolayer of 142 8+ was deposited on an<br />

array of microcantilever beams which had been coated on one<br />

side with a gold film <strong>and</strong> the set-up then inserted in a fluid<br />

cell. The chemical oxidant Fe(ClO 4) 3 was added to the<br />

solution <strong>and</strong> the combined effect of co-conformational<br />

change in approximately six billion r<strong>and</strong>omly oriented<br />

rotaxanes on each cantilever was an upward bending of the<br />

beam by about 35 nm. Reduction with ascorbic acid returned<br />

the cantilever to its starting position. The process could be<br />

repeated over several cycles,albeit with gradually decreasing<br />

amplitude. [509]<br />

8.3.4. Using <strong>Mechanical</strong> Switches To Affect Interfacial Properties<br />

The surface properties of an object are crucial in<br />

determining many aspects of its behavior <strong>and</strong> the ability to<br />

remotely change these characteristics could have a significantly<br />

impact on both current <strong>and</strong> future technologies. Many<br />

of the stimuli-switchable surfaces developed to date rely on<br />

induced submolecular motions to bring about the change in<br />

properties. [510]<br />

Anchoring stimuli-responsive polymers to solid substrates<br />

has allowed the creation of switchable surfaces based on<br />

Scheme 85. Chemical structure of [3]rotaxane 142 8+ <strong>and</strong> schematic representation of its operation as a molecular actuator. [508]<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

153


Reviews<br />

environmentally responsive<br />

hydrogel polymers <strong>and</strong><br />

polyelectrolytes,photoresponsiveazobenzene-containing<br />

polymers,<strong>and</strong> dendritic<br />

polymers,as well as<br />

changes arising from microphase<br />

segregation mechanisms<br />

for block-copolymer<br />

<strong>and</strong> mixed polymer brushes.<br />

These approaches have<br />

been extensively reviewed<br />

elsewhere, [510,511] <strong>and</strong> can<br />

give rise to switchable surface<br />

patterning,wettability,<br />

adhesion,<strong>and</strong> control of<br />

other interfacial interactions.<br />

Integration of these<br />

functional organic structures<br />

with inorganic <strong>and</strong><br />

biomaterials is an important<br />

current goal,particularly<br />

with a view to potential<br />

applications. [511d] For example,the<br />

intercalation of Au<br />

nanoparticles into an electrode-mountedsolventresponsive<br />

polymer gel<br />

allowed control of the interfacial<br />

conductivity through<br />

variation in the separation<br />

of the nanoparticles. [512] The<br />

encapsulation of a thermally<br />

responsive polymer in a porous inorganic matrix<br />

resulted in a switchable molecular filter, [513,514] while a<br />

covalently tethered monolayer of the same polymer has<br />

been used to reversibly adsorb proteins in an integrated<br />

microfluidics chip. [515]<br />

The direct deposition of bistable small molecules onto<br />

solid substrates as thin films can also result in interesting<br />

switching phenomena which amplify the molecular-level<br />

motion to give a large response. On polar surfaces such as<br />

mica,monolayer films of amphiphilic benzylic amide [2]catenane<br />

91 (Scheme 53) prefer to adopt co-conformation alkyl-<br />

91,irrespective of the nature of the depositing solution. [516]<br />

Continued deposition from acetone results in stacked monomolecular<br />

terraces that are clearly observable by AFM.<br />

Further deposition from chloroform,however,results in the<br />

incoming molecules (adopting the amido-91 form in the<br />

apolar environment) inducing a co-conformational change in<br />

the film so that amido-91 is adopted by all the molecules,thus<br />

minimizing the packing energy. The result is a completely<br />

different morphology: decreased interaction with the substrate<br />

causes complete rupture of the film to give stacks<br />

several tens of nanometers in height.<br />

A rotaxane-based molecular shuttle has been attached to<br />

mesoporous silica particles <strong>and</strong> used to control access to pores<br />

in the material (Figure 50)—a reversible analogue of an<br />

earlier pseudorotaxane-based design (Section 8.1). [517] When<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 50. a) Chemical structure of a bistable [2]rotaxane (143 4+ ) used to create a reversible molecular<br />

valve. [517] b) Schematic representation of the operating cycle for a reversible molecular valve. Colors not<br />

defined in (a) are: pink = TTF + ; light blue = guest molecules.<br />

the CBPQT 4+ ring sits on the preferred TTF station,access to<br />

the interior of the nanoparticles is unrestricted <strong>and</strong> diffusion<br />

of solutes from the surrounding solution can occur. Chemically<br />

induced oxidation of the TTF unit results in a shuttling<br />

of the ring closer to the solid surface,thereby blocking access<br />

to the pores <strong>and</strong> trapping any solute molecules inside.<br />

Reduction of the TTF unit reverses the mechanical motion,<br />

thus releasing the guest molecules <strong>and</strong> returning the system to<br />

its initial state.<br />

A potentially useful way of detecting <strong>and</strong> subsequently<br />

communicating the state of any type of molecular switch is to<br />

translate its state into an electrical signal. [97s,403e,518] This has<br />

been accomplished for a molecular shuttle attached through<br />

one end of the thread to the surface of a gold electrode. [519]<br />

Photochemically induced shuttling of a cyclodextrin ring<br />

closer to the electrode surface was detectable as an increase in<br />

the rate of electron transfer on oxidation of a redox-active<br />

ferrocene unit attached to the mobile ring. Shuttling has also<br />

been invoked to explain the operation of a remarkable device,<br />

in which a rotaxane connects the redox-active enzyme glucose<br />

oxidase to an electrode surface. [520] When the device is<br />

constructed without the rotaxane macrocycle (a CBPQT 4+<br />

cyclophane) no electronic communication between the redox<br />

enzyme <strong>and</strong> the electrode is detected. With the rotaxane<br />

linker,however,bioelectrocatalyzed oxidation of glucose<br />

occurs. It is proposed that the cyclophane acts as a two-<br />

154 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

electron relay,that is,acting as an intermediate for electron<br />

transfer between the enzyme <strong>and</strong> the electrode. It is likely<br />

that this role involves movement of the reduced ring towards<br />

the electrode as reduction destroys its charge-transfer interactions<br />

with the thread. Such a mechanism may also be at play<br />

in an earlier,fully synthetic,system where the ring acts as a<br />

relay between a photoelectroactive CdS nanoparticle <strong>and</strong> the<br />

electrode. [521] In a structurally simpler synthetic system<br />

(Scheme 86),electrochemically induced shuttling has been<br />

directly proven <strong>and</strong> shown to switch a number of properties of<br />

the functionalized electrode. [522] The cyclophane can be<br />

reversibly reduced <strong>and</strong> oxidized by applying appropriate<br />

potentials at the electrode. Chronoampometry demonstrates<br />

that the reductive electron transfer is significantly slower than<br />

the oxidation,thus indicating a shorter electrode–macrocycle<br />

separation in the reduced state. Positional integrity of the<br />

macrocycle in both states was quantitative to the limits of<br />

detection,however,fast two-potential-step chronoampometry<br />

experiments allowed a detailed study of the kinetics for<br />

shuttling (k 1 <strong>and</strong> k 2 in Scheme 86) in each direction at various<br />

temperatures <strong>and</strong> in solvent systems of differing viscosity. [522b]<br />

Impedance spectroscopy indicated a difference in the doublecharge-layer<br />

capacitance in the two states,while the decrease<br />

of the cyclophane charge <strong>and</strong> shuttling to reveal the alkyl<br />

chain-based thread also increases the hydrophobicity of the<br />

electrode surface. This was demonstrated by a change in the<br />

contact angles for an electrolyte droplet placed on the<br />

electrode. [522a]<br />

Inspired by a number of natural phenomena,the design of<br />

surfaces with special—<strong>and</strong> in particular switchable—wettability<br />

characteristics has attracted much interest in recent<br />

years. [523] Monolayers of various photochromic switches have<br />

been used to control the wettability of a surface. [510] In line<br />

Angew<strong>and</strong>te<br />

Chemie<br />

with an earlier computational study, [524] variation of an<br />

electrostatic potential has been used to alter the conformation<br />

of a SAM of long-chain alkanethiols terminated with<br />

carboxylate groups <strong>and</strong> thus affect the surface hydrophobicity.<br />

[525] Applying a negative potential to the gold surface repels<br />

the carboxylate groups,which causes the chains to “st<strong>and</strong> up”<br />

<strong>and</strong> thus endows a strongly hydrophilic character to the<br />

surface; a positive Au potential attracts the carboxylate<br />

groups,which causes folding that exposes the hydrophobic<br />

alkyl chains at the surface. A low-density SAM must be<br />

utilized to allow significant conformational freedom. Similar<br />

switchable surface-wettability properties were also demonstrated<br />

by using positively charged bipyridinium end groups<br />

instead of the negatively charged carboxylate groups. [526]<br />

More recently,a carboxylate-terminated low-density SAM<br />

was manufactured by self-assembling cyclodextrin pseudorotaxanes<br />

on a gold surface before washing off the spacefilling<br />

macrocycles with a polar solvent. [527] The resulting<br />

surface demonstrated conformational switching <strong>and</strong> could be<br />

used to reversibly adsorb both a cationic <strong>and</strong> a near-neutral<br />

protein at the interface.<br />

A dramatic illustration of the power of controlled<br />

molecular-level motion is the use of conformationally switchable<br />

monolayers to control macroscopic liquid transport<br />

across surfaces. It has been shown both theoretically [528] <strong>and</strong><br />

experimentally [529] that a liquid droplet on a surface with<br />

spatially inhomogeneous surface free energy will tend to<br />

move in the direction of the higher surface free energy. [530,531]<br />

Surface motions of droplets have been controlled by wettability<br />

gradients <strong>and</strong> steps in a number of different<br />

ways, [529,530,532,533] yet it remains a particular challenge to<br />

endow a surface with remotely controllable surface free<br />

energy so as to generate macroscopic motion of a droplet<br />

Scheme 86. Electrochemically induced shuttling on an electrode surface. [522] The submolecular co-conformational motion is transduced into an<br />

electronic signal in terms of a change in the electron-transfer kinetics <strong>and</strong> also results in a change in the capacitance <strong>and</strong> hydrophobicity of the<br />

monolayer. The rate constants shown for the shuttling steps (k 1 <strong>and</strong> k 2) are those measured at 258C in pure water. Only k 2 was found to vary with<br />

temperature, while both rates decreased with increasing solvent viscosity.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

155


Reviews<br />

along any path. [534] As well as addressing a number of<br />

fundamental scientific issues,such a system would undoubtedly<br />

be of major value in the emerging fields of microfluidics<br />

<strong>and</strong> lab-on-a-chip technologies which currently rely on<br />

potentially damaging strong electric fields,air bubbles,or<br />

expensive microscopic pumps to transport small quantities of<br />

liquids. [535]<br />

A monolayer of azobenzene-substituted calix[4]arene 144<br />

(Scheme 87) provides a photoresponsive surface on which the<br />

Scheme 87. <strong>Molecular</strong> structure of the azobenzene unit used to create<br />

a photoresponsive surface for the control of liquid motion. [536] The<br />

macrocyclic calix[4]arene scaffold ensures sufficient free volume to<br />

allow full configurational motion of the azobenzene moieties, even in a<br />

tightly packed monolayer.<br />

motion of liquid droplets could be controlled simply by<br />

irradiation with light. [536] A droplet of olive oil on the all-cis<br />

surface was irradiated at 436 nm with an asymmetrical<br />

intensity so that more isomerization to the trans form<br />

occurred on one side of the droplet compared to the other.<br />

The resulting gradient in the surface free energy caused the<br />

droplet to move away from the trans-rich region. This motion<br />

could be continued by “chasing” the droplet with the light,or<br />

else by resetting the surface at the front of the drop to a cisrich<br />

state with a spatially controlled irradiation at 365 nm,<br />

followed by a further cycle of 436 nm towards the rear <strong>and</strong> so<br />

on,with the motion carried out in a stepwise fashion. The<br />

potential utility of this approach was demonstrated in a<br />

number of ways: a liquid droplet could be used to move a<br />

millimeter-sized glass bead across the surface; the same<br />

approach could be used to functionalize a capillary tube <strong>and</strong><br />

move liquids along it; <strong>and</strong> finally,an organic chemical<br />

reaction could be performed by bringing two droplets<br />

together,each containing one of the reactants. [536]<br />

Recently,a similar effect was demonstrated using a<br />

photoresponsive surface based on molecular shuttles. [537]<br />

The millimeter-scale directional transport of diiodomethane<br />

drops across a surface (Figure 51) was achieved using the<br />

biased Brownian motion of the components of stimuliresponsive<br />

rotaxane 145 (Scheme 88) to expose or conceal<br />

fluoroalkane residues <strong>and</strong> thereby modify the surface tension.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 51. Lateral photographs of light-driven directional transport of a<br />

1.25 mL drop of diiodomethane across the surface of an (E)-145·11-<br />

MUA·Au(111) substrate on mica arranged flat (a)–(d) <strong>and</strong> up a 128<br />

incline (e)–(h). [537] a) Before irradiation (pristine (E)-145). b) After<br />

215 s of irradiation (20 s prior to transport) with UV light in the<br />

position shown (the right edge of the droplet <strong>and</strong> the adjacent<br />

surface). c) After 370 s of irradiation (just after transport). d) After<br />

580 s of irradiation (at the photostationary state). e) Before irradiation<br />

(pristine (E)-145). f) After 160 s of irradiation (just prior to transport)<br />

with UV light in the position shown (the right edge of the droplet <strong>and</strong><br />

the adjacent surface). g) After 245 s of irradiation (just after transport).<br />

h) After 640 s irradiation (at the photostationary state). For clarity, a<br />

yellow line is used on photographs (f)–(h) to indicate the surface of<br />

the substrate.<br />

Scheme 88. Stimuli-induced positional change of the macrocycle in<br />

the fluorinated molecular shuttle 145·2H + . [537]<br />

156 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

The collective operation of a monolayer of the molecular<br />

shuttles attached to a SAM of 11-mercaptoundecanoic acid<br />

(11-MUA) on Au(111) (Figure 52) was sufficient to power the<br />

movement of the microlitre droplet up a twelve degree incline<br />

Figure 52. A photoresponsive surface based on switchable fluorinated<br />

molecular shuttles. [537] Light-switchable rotaxanes with the fluoroalkane<br />

region (orange) exposed ((E)-145) were physisorbed onto a selfassembled<br />

monolayer of 11-MUA on Au(111) deposited onto either<br />

glass or mica to create a polarophobic surface, (E)-145·11-MUA·Au-<br />

(111). Illumination with light of wavelength 240–400 nm isomerizes<br />

some of the E olefins to Z causing a nanometer displacement of the<br />

rotaxane threads in the Z shuttles which encapsulate the fluoroalkane<br />

units, thus leaving a more polarophilic surface, (E)/(Z)-145·11-<br />

MUA·Au(111). The contact angles of droplets of a wide range of<br />

liquids change in response to the isomerization process.<br />

(Figure 51 e–h). In doing this,the molecular machines effectively<br />

employ the energy of the isomerizing photon to do<br />

work on the drop against gravity. In this experiment,<br />

approximately 50 % of the light energy absorbed by the<br />

rotaxanes was used to overcome the effect of gravity,with the<br />

work done stored as potential energy (the raised position of<br />

the droplet up the incline).<br />

The above examples demonstrate quite extraordinary<br />

application of collective submolecular motions to alter the<br />

macroscopic properties of a surface <strong>and</strong> in turn induce<br />

movement of a macroscopic object. [538] However,the changes<br />

in the surface free energy involved are relatively small. It is<br />

well known that surface roughness can amplify both hydrophobicity<br />

<strong>and</strong> hydrophilicity. [523] A micropatterned surface<br />

functionalized with a solvent-sensitive mixed polymer brush<br />

could switch between a hydrophilic <strong>and</strong> a superhydrophobic<br />

state. [539] Reversible switching between superhydrophobicity<br />

<strong>and</strong> superhydrophilicity has been achieved on a static micropatterned<br />

surface, [540] as well as on a dynamic nanostructured<br />

surface, [541] in both cases covered with a temperature-responsive<br />

polymer. A similar approach has been used to amplify the<br />

light-induced wettability switching for a spiropyran-functionalized<br />

surface. [542] As well as increasing the change in the<br />

contact angle relative to the smooth surface on irradiation,<br />

contact angle hysteresis was also reduced on the rough surface<br />

so that a water droplet could be moved across it on<br />

application of an asymmetric irradiation.<br />

9. Artificial Biomolecular <strong>Machines</strong><br />

9.1. Hybrid Biomolecular <strong>Motors</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

Motor proteins provide both inspiration <strong>and</strong> important<br />

proofs-of-principle for the synthetic chemist trying to create<br />

artificial motor molecules. In turn,the increasing sophistication<br />

of the task carried out by synthetic structures may one<br />

day (perhaps sooner than one might think) start to aid our<br />

underst<strong>and</strong>ing of complex natural systems. However,the<br />

demonstration of such amazing mechanical functionality by<br />

motor proteins has also triggered a different approach to<br />

creating tiny mechanical devices: namely the direct application<br />

of complete biomolecular motor constructs in nonnatural<br />

settings. Initially,interfacing operational motor proteins<br />

with artificial constructs was purely concerned with<br />

further dissecting the mechanochemical mechanisms of the<br />

natural systems. In a series of experiments which have<br />

achieved classic status,direct observation of directional<br />

rotation of the F1-ATPase motor,driven by ATP hydrolysis,<br />

was realized for individual motors mounted on glass, [543] while<br />

the transmission of this motion to the F0 motor in complete<br />

F0F1-ATPase constructs has also been observed. [544] The<br />

reverse process—ATP synthesis—has subsequently been<br />

achieved in a similar environment,by physically driving<br />

rotation of the F1 motor in the opposite direction to that<br />

observed for the hydrolysis reaction; this is the first demonstration<br />

of artificial chemical synthesis driven by a vectorial<br />

force! [545]<br />

Mimicry <strong>and</strong> deconvolution of natural systems aside,<br />

methods by which biomolecular motors can be reliably<br />

incorporated into artificial systems are being pursued to<br />

create nanomechanical systems powered by ready-made<br />

molecular motors. [546] Such endeavors constitute a research<br />

area in their own right, [547]<br />

involving,for example,the<br />

integration of soft organic functional molecules with hard<br />

inorganic components <strong>and</strong> supports; the genetic engineering<br />

of motor proteins to adapt their function <strong>and</strong> interaction with<br />

synthetic components; <strong>and</strong> the engineering of switches<br />

whereby the natural motor can be turned on or off.<br />

The use of myosin molecules immobilized on a solid<br />

substrate to transport actin str<strong>and</strong>s across the surface was one<br />

of the earliest in vitro studies of motor protein mechanochemistry.<br />

[548] Analogous systems,many employing kinesin<br />

(or dynein) <strong>and</strong> microtubules as the motor <strong>and</strong> cargo<br />

components,respectively,are now being investigated with a<br />

view to practical applications. [549] The potential uses of such<br />

systems clearly include the transport of nanoscale cargoes <strong>and</strong><br />

consequently the challenges of controlling <strong>and</strong> directing the<br />

motion,as well as capture <strong>and</strong> release of molecular materials,<br />

are actively being pursued. [1l,547a,c] As mentioned in Section<br />

8.3.3,temperature-responsive polymers can be used as<br />

control elements for motor proteins, [476] but recently it was<br />

shown that proteins themselves can be turned into novel<br />

polymer gel actuators. In particular,a chemically cross-linked<br />

actin gel has been shown to move over a cross-linked myosin<br />

gel in the presence of ATP,with speeds achieved similar to<br />

those of the native protein construct,thus opening up the<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

157


Reviews<br />

possibility of creating gel-based devices powered by protein<br />

motors rather than by osmotic effects. [550]<br />

As for fully synthetic systems,any macroscopic mechanical<br />

application of biomolecular motors is likely to require the<br />

parallel operation of a great number of molecular devices—<br />

just as limb movement in an animal is the result of the<br />

coordinated effort of many millions of myosin motors.<br />

Recently,the self-assembly of muscle cells on silicon microdevices<br />

has been achieved. The advantage of using whole cells<br />

(cardiomyocytes in this case) to grow muscle bundles is that<br />

all the machinery for cooperative motion is automatically<br />

present <strong>and</strong> the result is the first microstructure which can<br />

move autonomously as a result of cooperative contraction of<br />

muscle bundles. [551]<br />

A related active field is that of biomolecular electronics,in<br />

which native <strong>and</strong> genetically engineered proteins are<br />

deployed in molecular-based electronic <strong>and</strong> photonic devices.<br />

In many cases the electronic switching mechanism involves<br />

molecular-level motion,exemplified in particular by systems<br />

which rely on the photochromic <strong>and</strong> photoelectric properties<br />

of bacteriorhodopsin. [552]<br />

9.2. Hybrid Membrane-Bound <strong>Machines</strong><br />

The development of synthetic ionophores <strong>and</strong> ion channels<br />

has been a central goal of supramolecular chemistry for a<br />

number of decades <strong>and</strong> advanced functionality such as<br />

enantioselection <strong>and</strong> stimuli-controlled transport has been<br />

achieved. [553] Recently,however,biological channel proteins<br />

have been used in artificial settings <strong>and</strong> external control over<br />

their functions achieved. The mechanosensitive channel of<br />

large conductance (MscL) from Escherichia coli is a homopentameric<br />

protein channel which opens in response to<br />

internal pressure within the bacterium,thus allowing nonselective<br />

efflux of ions <strong>and</strong> small solutes. [554] It is known,<br />

however,that the introduction of polar or charged residues at<br />

a specific location in the protein can result in spontaneous<br />

opening of the pore. [555] Incorporation of a cysteine residue at<br />

this crucial amino acid position provided a h<strong>and</strong>le to which<br />

Feringa <strong>and</strong> co-workers could attach artificial photoswitches,<br />

thereby creating light-switchable versions of this system. [556]<br />

In the first instance,an acetate unit protected by a photocleavable<br />

group was attached to the cysteine residue (Scheme<br />

89a) <strong>and</strong> the resulting hybrid (146) incorporated in a synthetic<br />

lipid membrane. Photolysis of the protecting group revealed<br />

free acetate units,<strong>and</strong> concomitant opening of the channels<br />

was clearly observed on the single-molecule level by using<br />

patch-clamp techniques. A reversible analogue was created<br />

by appending a spiropyran switch to the cysteine residue<br />

(giving SP-147,Scheme 89b). Opening <strong>and</strong> closing of the<br />

channel was observed again in patch-clamp experiments,<br />

however,a significant decrease in the number of pores<br />

opening was observed on repeating the procedure for a<br />

second time. An efflux experiment was performed in which a<br />

self-quenching dye was released from liposomes on opening<br />

the channel. A clear increase in permeability on irradiation<br />

was demonstrated,although some leakage from the “closed”<br />

channels was also observed. A related approach has<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Scheme 89. Chemical structures of the photochromic switches used to<br />

create: a) irreversible <strong>and</strong> b) reversible light-triggered MscL. [556] As the<br />

MscL construct is a homopentamer, each channel features five sites<br />

for attachment of the photochromic units. SP = spiropyran form,<br />

MC = merocyanine form.<br />

employed azobenzene chromophores to reversibly place a<br />

blocking group in Shaker K + ion channels in rat hippocampal<br />

neurons [557] <strong>and</strong> to bring an agonist into contact with the<br />

binding site in a lig<strong>and</strong>-gated ion channel,again in living<br />

cells. [558] The latter could,in principle,be a general method for<br />

creating optically switchable versions of many allosterically<br />

regulated proteins. Rather than postsynthetic functionalization<br />

of amino acid side chains,photoswitches can also be<br />

incorporated by chemical synthesis as non-natural amino<br />

acids in place of terminal residues. This approach has been<br />

used to create various photoswitchable analogues of the lowmolecular-weight<br />

channel-forming peptide gramicidin. [559]<br />

Artificial carrier mediated translocation of ions across a<br />

membrane can be coupled to redox reactions so that the<br />

carrier is oxidized <strong>and</strong> reduced at opposite sides of the<br />

membrane,thus adjusting its affinity for the transported<br />

species depending on its location. [560] A particularly impressive<br />

extension of this concept draws on the mechanism of<br />

photosynthetic energy conversion in nature—the photogeneration<br />

of charge-separated states. [561] Artificial photosynthetic<br />

reaction center 148 (Scheme 90) exploits components<br />

commonly used by nature—a porphyrin photosensitizer (P),a<br />

carotenoid polyene electron donor (D),<strong>and</strong> a naphthoquinone<br />

electron acceptor—to create charge-separated states of<br />

the form D + C-P-A C. Directionally inserting 148 into liposome<br />

membranes results in a transmembrane electrochemical<br />

potential on irradiation with light. This allows the spatial<br />

control of the oxidation state for quinone-based carrier<br />

molecules which are dissolved <strong>and</strong> freely diffusing in the<br />

membrane. The change in oxidation state switches the<br />

pK a value [562] or the calcium ion binding affinity [563] of the<br />

carrier molecules,depending on their proximity to the inner<br />

or outer surface of the membrane,so as to facilitate proton or<br />

calcium ion pumping across these “artificial photosynthetic<br />

membranes”. Continual operation of the proton transport<br />

system <strong>and</strong> concomitant generation of a protonmotive force<br />

158 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Scheme 90. Artificial photosynthetic reaction center 148 which was used to create a transmembrane electric potential so as to power proton [562,564]<br />

or calcium [563] transport.<br />

was illustrated by incorporation of intact F 0F 1-ATPase into<br />

the artificial photosynthetic membrane,<strong>and</strong> in vitro ATP<br />

synthesis was successfully demonstrated. [564]<br />

These systems harness directional electron transfer to set<br />

up the electrochemical gradient necessary for ion transport.<br />

They therefore correspond to the redox-loop (or Mitchellloop)<br />

mechanisms for transport in nature (Section 1.3). [33]<br />

<strong>Synthetic</strong> examples of conformational pumps,on the other<br />

h<strong>and</strong>,have not yet been realized,although the design of<br />

potential components is being actively pursued. [565]<br />

9.3. DNA-Based Switches <strong>and</strong> <strong>Motors</strong>: Molecules that Can Walk<br />

It is also possible to borrow from nature,not the<br />

completed machine or some components thereof,but the<br />

materials used to build them. The use of nucleic acids as<br />

nanoscale construction materials have been exquisitely demonstrated<br />

over the past decade in a variety of ways. [566,567]<br />

Following an initial report on the control of double-str<strong>and</strong>ed<br />

DNA branch migration, [568] the first artificial,well-defined,<br />

large amplitude DNA-based conformational switch was<br />

achieved by harnessing the ability of certain DNA sequences<br />

to change between right-h<strong>and</strong>ed B forms <strong>and</strong> left-h<strong>and</strong>ed<br />

Z forms,depending on the nature of the environment. [569] One<br />

such sequence was used to connect two rigid DNA doublecrossover<br />

motifs to create a linear mechanical device. The B–<br />

Z transition results in a twisting of the device around the<br />

central unit (as well as a small change in length),thereby<br />

altering the separation of equivalent points on the two<br />

double-crossover units. This reversible process was monitored<br />

by fluorescence resonance energy transfer (FRET) between<br />

dyes attached to the double-crossover portions. [569]<br />

The system depicted in Figure 53 is not only constructed<br />

from DNA but is also operated by the addition of specific<br />

oligonucleotide sequences. Three DNA str<strong>and</strong>s (A, B,<strong>and</strong> C)<br />

self-assemble to give the “relaxed” structure. A fourth str<strong>and</strong>,<br />

the “fuel” F,has sequences complementary to the singlestr<strong>and</strong>ed<br />

regions of both A (red) <strong>and</strong> B (blue),<strong>and</strong> so<br />

hybridization occurs to give the “closed” form. However, F<br />

also contains an eight-base overhang (pink) at its 3’ end. This<br />

acts as a “toehold” for a “removal” str<strong>and</strong> R,which is<br />

complementary to the full length of F. Binding at the<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Figure 53. Operation of a pair of DNA tweezers. [570] Closing of the<br />

tweezers amounts to bringing the two rigid double-str<strong>and</strong>ed regions<br />

(black) close together by hybridization of a fuel str<strong>and</strong> F. Removal of<br />

the fuel str<strong>and</strong> to restore the relaxed state is brought about by addition<br />

of a removal str<strong>and</strong> R, which binds initially at the toehold on F, then<br />

branch migration leads to complete removal of F from the device with<br />

concomitant formation of one molecule of duplex waste FR. The<br />

coloring scheme indicates the complementary regions on different<br />

threads; lines indicating base paring are purely schematic <strong>and</strong> do not<br />

represent a particular number of bases. TET = 5’-tetrachlorofluorescein<br />

phosphoramidite; TAMRA = carboxytetramethylrhodamine.<br />

overhang is followed by continuing hybridization along the<br />

length of the two str<strong>and</strong>s by a branch-migration process,until<br />

the inert “waste” duplex FR is formed <strong>and</strong> the machine<br />

returns to its original relaxed state. The process is reversible<br />

<strong>and</strong> reproducible over several cycles,as evidenced by FRET<br />

measurements which indicate a difference in the end–end<br />

distance between the two forms of approximately 6 nm. [570] A<br />

related device adopts the same strategy to switch between a<br />

relaxed state <strong>and</strong> an “open” state,where addition of the fuel<br />

str<strong>and</strong> results in an approximately linear conformation. [571]<br />

Finally,these two systems were combined to create a threestate<br />

device capable of switching repeatedly between conformationally<br />

well-defined open <strong>and</strong> closed forms through the<br />

more flexible relaxed state. [572] Switching between two welldefined<br />

conformations triggered by hybridization has also<br />

been achieved in an alternative system based on crossover<br />

motifs. [573] In this case,a 1808 rotation of one end of a linear<br />

device relative to the other end results. One-dimensional<br />

oligomeric arrays of such devices were prepared,with a<br />

structurally rigid DNA substituent appended to each monomer.<br />

The switching motion could then be observed by AFM as<br />

a change in the relative orientation of the rigid subunits. A<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

159


Reviews<br />

related system has since been designed to exhibit lengthening<br />

<strong>and</strong> contraction of a linear DNA construct. [574] This device was<br />

incorporated into a 2D DNA lattice <strong>and</strong> its switching used to<br />

reversibly alter the dimensions of the lattice features,again<br />

evidenced by AFM.<br />

Opening <strong>and</strong> closing or twisting of DNA devices could<br />

clearly be used to alter the relative orientations of a wide<br />

number of attached functionalities,in a manner similar to<br />

rotaxane-based co-conformational switches (Section 8.2.2),<br />

while it has also been postulated that these devices may be<br />

able to exert mechanical forces,again similar to the shuttlebased<br />

switches (Section 8.3.3). The use of specific oligonucleotides<br />

to trigger the molecular-level motion may have<br />

advantages for the development of more complex systems,<br />

where it should be possible to construct machines composed<br />

of a variety of subtly different molecular components. Each<br />

could then be selectively triggered by a specific oligonucleotide<br />

stimulus,allowing the fuel to act also as an information<br />

carrier between the operator <strong>and</strong> machine,or even between<br />

different machine components. On the downside,however,<br />

each operational cycle results in the production of a waste<br />

duplex str<strong>and</strong> <strong>and</strong>,from a practical point of view,dilution of<br />

the system. Both these factors have been shown to result in a<br />

decrease in the fluorescent response on increasing cycles.<br />

Furthermore,the hybridization processes are relatively slow:<br />

half-times for completion of these reactions are commonly of<br />

the order of tens of seconds. On the other h<strong>and</strong>,a DNA<br />

conformational switch which operates simply by adding <strong>and</strong><br />

removing protons has now been created (Figure 54). [575]<br />

Certain cytosine-rich str<strong>and</strong>s become partially protonated at<br />

low pH values <strong>and</strong> form a compact quadruplex structure<br />

known as the i motif. Raising the pH value disrupts this<br />

structure,thereby allowing the single-str<strong>and</strong>ed DNA to be<br />

captured by a partially complementary str<strong>and</strong> waiting in<br />

solution,thus forming an extended duplex structure,as<br />

evidenced by a decrease in the FRET response between two<br />

attached dyes. If the binding between the two str<strong>and</strong>s is not<br />

too strong,the generation of an acidic environment reforms<br />

the i motif. The system exhibits impressive fatigue resistance<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

even after 30 cycles,while the switching processes both occur<br />

in about 5 s. [576] Two related triplex–duplex contraction–<br />

expansion systems,also triggered by controlling cytosine<br />

protonation,have also been reported. [577] Both are fully selfcontained<br />

devices in which all the oligonucleotides (two<br />

st<strong>and</strong>s in one case,three str<strong>and</strong>s in the other) remain<br />

associated in both open <strong>and</strong> closed forms. Quadruplex–<br />

duplex contraction–expansion systems have also been achieved<br />

by using the competitive hybridization strategies<br />

discussed above, [578] <strong>and</strong> this has been combined with aptamer<br />

technology to create a device in which binding <strong>and</strong> release of<br />

a protein (the human blood clotting factor a-thrombin) can be<br />

reversibly achieved. [579] It is worth noting that these contraction–expansion<br />

systems bear some resemblance to the switchable<br />

helical oligomeric systems discussed in Section 2.1.4.<br />

In most of the switches <strong>and</strong> motors we have discussed so<br />

far,at least two stimuli,applied consecutively,are required to<br />

send the machine through an operating cycle. A DNA-based<br />

open–closed conformational switch has been created,however,which<br />

can continuously cycle between two states<br />

without any external intervention as long as its fuel is<br />

present. [580] The two-str<strong>and</strong>ed cyclic structure (Figure 55)<br />

consists of two duplex arms connected by a single base<br />

“hinge” at one end <strong>and</strong> a longer single-str<strong>and</strong>ed region at the<br />

other. The 29-base single-str<strong>and</strong>ed region (E) contains an<br />

RNA-cleaving 10–23 DNA enzyme (light blue) <strong>and</strong> in the<br />

absence of substrate (<strong>and</strong> in the presence of divalent cations)<br />

Figure 54. pH-switched quadruplex–duplex switching in a DNA molecular<br />

machine. [575] In acidic environments, the cytosine-rich str<strong>and</strong> A is<br />

partially protonated <strong>and</strong> forms the four-str<strong>and</strong>ed self-complementary<br />

i motif. This holds the two dyes in proximity <strong>and</strong> FRET occurs.<br />

Increasing the pH value turns off the self-complementary interactions<br />

(blue!red) <strong>and</strong> the str<strong>and</strong> can now be captured by a partially<br />

complementary partner (B), thereby forming an extended duplex<br />

structure which holds the fluorophore <strong>and</strong> quencher units far apart<br />

<strong>and</strong> reducing the FRET response. The coloring scheme indicates the<br />

complementary regions on different threads; lines indicating base<br />

paring are purely schematic <strong>and</strong> do not represent a particular number<br />

of bases. Figure 55. An autonomous DNA machine. [580, 581] The machine consists<br />

of two single str<strong>and</strong>s (D <strong>and</strong> E). The single-str<strong>and</strong>ed 10–23 DNAzyme<br />

region (light blue) of E adopts a coiled-coil conformation so that the<br />

machine adopts a closed arrangement. Binding of a DNA–RNA<br />

chimera substrate (S) forces the machine to adopt an open state.<br />

Subsequent enzymatic action on the substrate creates two products<br />

(S 1 <strong>and</strong> S 2 ), each of which only has weak complementarity for the<br />

enzyme str<strong>and</strong> <strong>and</strong> so are released in favor of the closed conformation.<br />

Addition of an all-DNA str<strong>and</strong> B which is complementary to the activesite<br />

region, also forces formation of the open conformation, but halts<br />

enzyme action until B is removed by competitive hybridization using<br />

removal str<strong>and</strong> R. The coloring scheme indicates the complementary<br />

regions on different threads; lines indicating base paring are purely<br />

schematic <strong>and</strong> do not represent a particular number of bases.<br />

160 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

it exists in a coiled-coil conformation which gives an overall<br />

closed state for the device. Substrates (S) for the DNAzyme<br />

are DNA–RNA chimeras complementary to regions either<br />

side of the catalytic site so that hybridization forces the<br />

structure to adopt an open arrangement. Cleavage of the<br />

substrate ensues <strong>and</strong> produces two fragments (S 1 <strong>and</strong> S 2 ),<br />

neither of which bind strongly to E <strong>and</strong> so are ejected in favor<br />

of forming the compact closed state once more. The device is<br />

now ready to undergo a second cycle of opening <strong>and</strong> closing<br />

<strong>and</strong> will continue to do so,without any external control,until<br />

no substrate “fuel” is left.<br />

Also illustrated in Figure 55 is an external control<br />

mechanism whereby the autonomous machine can be<br />

switched on or off by addition or removal of a “brake”<br />

str<strong>and</strong> (B),which has greater affinity for E than the substrate<br />

but cannot itself be cleaved. [581] A different autonomous<br />

device has been realized based on the elegant use of topology<br />

to kinetically hinder a competitive hybridization reaction. [582]<br />

In this case,the “device” is simply a short oligonucleotide<br />

which oscillates between a r<strong>and</strong>om single-str<strong>and</strong>ed state <strong>and</strong> a<br />

more-ordered duplex state.<br />

As with the small-molecule organic systems,a key<br />

challenge for DNA-based machines is achieving directional<br />

processive motion,either in a linear or rotary form. In this<br />

regard,successful systems have recently been demonstrated<br />

by four independent research groups. [583] The first of these,<br />

from Sherman <strong>and</strong> Seeman,is shown in Figure 56. [584] A triplecrossover<br />

molecule acts as a rigid track from which protrudes<br />

Figure 56. A non-autonomous DNA walker which moves using an<br />

inchworm-like gait. [584] Matching colors indicate complementary<br />

sequences between the str<strong>and</strong>s; lines indicating base paring are purely<br />

schematic <strong>and</strong> do not represent a particular number of bases. a) The<br />

walker starts off attached to the track through both “feet”, by means of<br />

“set” str<strong>and</strong>s which each have regions complementary to the appropriate<br />

foot <strong>and</strong> “foothold” on the track. b) The front foot is released by<br />

removal of the appropriate set str<strong>and</strong> through competitive hybridization,<br />

initiated at a single-str<strong>and</strong>ed toehold region (pink) on the set<br />

str<strong>and</strong>. The removal str<strong>and</strong> has biotin attached so that the resultant<br />

duplex waste can be easily removed. c) The front foot is attached to<br />

the next foothold on the track using a set str<strong>and</strong> of appropriate<br />

sequence. d) The rear foot is now released by competitive hybridization<br />

of its set str<strong>and</strong>. e) Finally, the rear foot is attached at its new<br />

position by the required set str<strong>and</strong>.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

a series of three single-str<strong>and</strong>ed “footholds” (red,green,<strong>and</strong><br />

light blue). A bipedal walking device is constructed from two<br />

helical domains (“legs”),each with a single-str<strong>and</strong>ed “foot”<br />

(dark blue <strong>and</strong> orange) at one end <strong>and</strong> connected together by<br />

flexible linkers at the other. Attachment of the walker to the<br />

track occurs through “set” str<strong>and</strong>s which have a specific footcomplementary<br />

sequence adjacent to a specific footholdcomplementary<br />

sequence. Starting with both feet anchored to<br />

the track (Figure 56 a),the “front” foot can be lifted by<br />

removing its set str<strong>and</strong> (Figure 56b). This is done by<br />

competitive hybridization initiated at a toehold region<br />

(pink) as discussed for the simpler machines above. Addition<br />

of a different set str<strong>and</strong> then anchors this foot to the next<br />

foothold (Figure 56 c). Finally,the rear foot is freed (Figure<br />

56d) <strong>and</strong> attached to the foothold just vacated by the<br />

front foot (Figure 56e). The walker can be returned to the<br />

starting position by a similar series of steps. Such a process,<br />

with the rear leg always trailing,has been described as an<br />

“inchworm” gait (see Section 4.4). The present system only<br />

contains three footholds <strong>and</strong> so there is no choice of direction<br />

for the walker to make at any stage; it is a positional switch<br />

rather than a motor. However,it may be possible to generate<br />

continued directional motion simply with a repeating<br />

sequence of three attachment points. It must be noted that<br />

the nature of the flexible linkers between the two legs is<br />

crucial here: they must be able to stretch across,say,the green<br />

foothold to link the red <strong>and</strong> light blue sites,but not so flexible<br />

as to allow motion in the wrong direction through passing of<br />

the front foot to the rear <strong>and</strong> attachment at a foothold on this<br />

side.<br />

A similar strategy was used by Shin <strong>and</strong> Pierce to create a<br />

bipedal walker (Figure 57) that moves with a gait in which the<br />

rear leg advances to the front (a “passing-leg” gait,see<br />

Figure 57. A non-autonomous DNA walker with a “passing-leg”<br />

gait. [585] Matching colors indicate complementary sequences between<br />

str<strong>and</strong>s; lines indicating base paring are purely schematic <strong>and</strong> do not<br />

represent a particular number of bases. A sequence of removal str<strong>and</strong><br />

<strong>and</strong> set str<strong>and</strong> additions removes the rear foot <strong>and</strong> reattaches it at the<br />

next foothold in front of the walker (see text for details). Each foothold<br />

is appended with a different fluorophore (not shown) <strong>and</strong> each foot a<br />

quencher, so that progression of the walker can be monitored by<br />

multiplexed fluorescence quenching measurements.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

161


Reviews<br />

Section 4.4). [585] The track was constructed from six oligonucleotides<br />

which form a helical scaffold from which four 20base<br />

single-str<strong>and</strong>ed footholds (red,green,light-blue,<strong>and</strong><br />

yellow) protrude at regular intervals. The walker consists of<br />

two partially complementary oligonucleotides which form a<br />

20-base helix at one end,leaving two 23-base single-str<strong>and</strong>ed<br />

legs (dark blue <strong>and</strong> orange). A sequence of set-str<strong>and</strong>mediated<br />

attachment <strong>and</strong> removal by competitive hybridization<br />

results in unidirectional walking of the device,as<br />

illustrated in Figure 57.<br />

A pair of DNA-based “gears” which can rotate unidirectionally<br />

against each other has been created by Ye <strong>and</strong> Mao<br />

by using the same principles (Figure 58). [586] This system<br />

comprises two DNA duplex circles composed of a central<br />

cyclic str<strong>and</strong> surrounded by three different linear str<strong>and</strong>s,<br />

each of which has a single-str<strong>and</strong>ed “tooth” at one end. The<br />

two gears can be connected through one tooth on each by a<br />

suitable set str<strong>and</strong>. The addition of a second set str<strong>and</strong> can<br />

attach the gears between two of the remaining teeth. Removal<br />

of the original set str<strong>and</strong> leaves a single linkage between the<br />

two gears again,but rotated by 1208 with respect to the<br />

starting point. The process can be continued in either<br />

direction to generate a full 3608 rotation.<br />

A number of approaches for achieving an autonomously<br />

processive DNA-based device have been proposed, [587] <strong>and</strong><br />

the first such system was recently reported by Turberfield <strong>and</strong><br />

Figure 58. Unidirectional correlated rotation of a pair of DNA-based gears, driven by<br />

hybridization reactions. [586] Matching colors indicate complementary sequences<br />

between str<strong>and</strong>s; lines indicating base paring are purely schematic <strong>and</strong> do not<br />

represent a particular number of bases.<br />

co-workers. [588] In this case,the walker is a 6-base fragment of<br />

DNA (colored red in Figure 59) which moves along three<br />

footholds on a track through a series of ligation <strong>and</strong><br />

restriction steps catalyzed by a ligase <strong>and</strong> two different<br />

restriction endonucleases. In the initial state the walker is<br />

ligated to foothold A. However,it has a 3-base sticky end<br />

which is complementary to the single-str<strong>and</strong>ed overhang of<br />

foothold B. The flexible single-str<strong>and</strong>ed regions at the base of<br />

each foothold allows hybridization of the complementary<br />

components (Figure 59b),thus creating a substrate for T4<br />

ligase <strong>and</strong> resulting in covalent attachment of the two DNA<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Figure 59. A unidirectional autonomous DNA walker. [588] The walker<br />

consists of the six-nucleotide base sequence colored red. Covalent<br />

attachment of the walker to a foothold is indicated by an asterisk;<br />

bridging between two footholds by hybridization is indicated by a plus<br />

sign; covalent attachment of the walker between two footholds is<br />

indicated thus: X*Y. In practice, the sequences of footholds A <strong>and</strong> C<br />

are identical. Lines indicating base paring are purely schematic <strong>and</strong> do<br />

not represent a particular number of bases.<br />

fragments (giving A*B,Figure 59c). In turn,this creates a<br />

recognition site for a restriction enzyme (PflM 1). Crucially,the<br />

selectivity of the resulting cleavage is such that<br />

the walker is always transferred onto foothold B (Figure<br />

59d) <strong>and</strong> the restriction site is destroyed. This means<br />

that although B* <strong>and</strong> A may associate again through<br />

hybridization (giving A + B*,Figure 59 e),<strong>and</strong> ligation<br />

may occur to give A*B once more,this is simply an<br />

“idling” step as subsequent cleavage can only restore B*<br />

once more: the motion is essentially ratcheted by the<br />

selectivity of the restriction endonuclease. Foothold C has<br />

the same overhanging three bases as foothold A,however,<br />

<strong>and</strong> so hybridization in this “forward” direction may occur<br />

(giving B* + C,Figure 59 f). The complex formed is again<br />

a substrate for T4 ligase,so B*C is produced (Figure 59 g).<br />

This creates a substrate for a second restriction endonuclease,BstAP<br />

1,<strong>and</strong> cleavage occurs to give C* (Figure<br />

59h). Again,an idling step involving religation with B<br />

may occur,but cleavage to give B* does not follow <strong>and</strong> so<br />

the motion cannot be reversed. This device then is an<br />

interesting example of an information ratchet (see Section<br />

1.4.2). The free energy of the device at states A*, B*,<br />

or C* is virtually identical <strong>and</strong> invariant,<strong>and</strong> the directionality<br />

of motion is achieved irrespective of what foothold is<br />

occupied,with the kinetics for passage in one direction (left<br />

to right as drawn) being faster than passage in the opposite<br />

direction.<br />

Two further autonomous walking systems have also been<br />

reported. [589,590] Both these machines involve tracks which<br />

display essentially identical footholds for the walker,together<br />

with a single enzyme which is able to cleave portions of the<br />

162 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

track only when the walker is bound. In one case, [589] a nicking<br />

enzyme is employed (Figure 60). This is a restriction endonuclease<br />

which binds double-str<strong>and</strong>ed DNA at a specific<br />

sequence but then catalyzes the cleavage of only one str<strong>and</strong>.<br />

Figure 60. An autonomous DNA walker (red) starting midway through<br />

its journey from left to right. [589] Lines indicating base paring are purely<br />

schematic <strong>and</strong> do not represent a particular number of bases.<br />

Hybridization of the walker oligonucleotide to a singlestr<strong>and</strong>ed<br />

foothold on the track provides the correct recognition<br />

sequence for the enzyme (Figure 60a),which subsequently<br />

cleaves the foothold str<strong>and</strong> eight bases from its end<br />

(Figure 60 b). The resultant eight-base duplex has a melting<br />

temperature below the operating temperature of the device,<br />

so dissociation occurs (Figure 60c) to reveal an overhang on<br />

the walker through which it can seek (Figure 60 d)—<strong>and</strong> then<br />

bind to—the next foothold along the track (Figure 60 d!e).<br />

Motion continues in this way,with movement in the “backwards”<br />

direction prevented by the irreversible cleavage of<br />

each successive foothold. The walker can be stopped on any<br />

foothold by introducing a single mismatch in the enzyme<br />

recognition site. In the second example, [590] the walker itself is<br />

the enzyme—a 10–23 DNAzyme,similar to that used in the<br />

autonomous tweezers (Figure 55). The footholds are provided<br />

by DNA–RNA chimeras <strong>and</strong> duplex formation with the<br />

walker results directly in cleavage of the foothold between<br />

two RNA residues. In a manner similar to the previous<br />

example,this frees a portion of the walker to seek out the next<br />

foothold with which it can fully associate. Just like the<br />

autonomous walker in Figure 59,both these systems operate<br />

by imposing an almost infinite barrier to motion in the<br />

“backwards” direction. Unlike the first example,however,<br />

these systems completely destroy the track in their wake—the<br />

motion is entirely irreversible <strong>and</strong> the track not reusable—<br />

while waste oligonucleotides are produced in each cleavage<br />

step. [591]<br />

The rapid progress in DNA-based molecular machines is<br />

testament to the power of this precisely programmable selfassembling<br />

construction <strong>and</strong> information-bearing material.<br />

[592] In living systems,oligonucleotides carry out a variety<br />

of roles within highly complex environments <strong>and</strong> the next<br />

advances in the artificial systems will surely see integration of<br />

these relatively simple machines,which exploit the structural<br />

properties of the material,with other emerging nucleotidebased<br />

technologies. Already a DNA tweezer device,closely<br />

related to that depicted in Figure 53,has been integrated with<br />

genetic DNA. [593] In this system,the genetic material is<br />

transcribed <strong>and</strong> produces a str<strong>and</strong> of mRNA which acts as the<br />

fuel str<strong>and</strong>,thereby closing the tweezers in an autonomous<br />

fashion. A similar process could be used to generate the<br />

removal str<strong>and</strong>,while mechanisms for regulating gene transcription<br />

have also been incorporated. The use of DNA as a<br />

templating scaffold for carrying out covalent chemistry is<br />

another emerging area where DNA switching technology has<br />

already been applied. [594] Here,a pH-controlled duplex–<br />

triplex switch is able to swap the reactivity of two chemically<br />

similar amines in an acylation reaction with a carboxylic acid.<br />

Furthermore,the machines described above exploit only a<br />

fraction of the structural characteristics <strong>and</strong> abilities of DNA,<br />

while DNA-templated assembly <strong>and</strong> nanofabrication of other<br />

materials is also of great interest. [566a,f–h,592,595] The recognition<br />

<strong>and</strong> mechanical properties of DNA have recently been<br />

exploited in a number of sensor systems. In particular,<br />

fluorophore-labeled DNA hairpins—so-called “molecular<br />

beacons”—are capable of very sensitive <strong>and</strong> specific detection<br />

of oligonucleotides. [596] The ability of certain DNA<br />

sequences to bind specific proteins,often resulting in a<br />

conformational change,has been exploited in a DNA-based<br />

device which can estimate binding free energies through a<br />

nanomechanical mechanism. [597] In another example,a DNA<br />

aptamer was combined with a fluorescently labeled,complementary<br />

oligonucleotide str<strong>and</strong> so that binding of the aptamer<br />

to its target (adenosine,in this case) is signaled by release of<br />

the tagged str<strong>and</strong>. [598] The released str<strong>and</strong> can then be<br />

designed to hybridize with another oligonucleotide in<br />

response to the analyte,while introduction of an enzyme<br />

that reacts with the guest molecule can reset the system.<br />

Indeed,the ability of oligonucleotides to selectively recognize<br />

specific sequences amongst a milieu of other molecular<br />

“noise”,together with the highly refined enzymatic tools of<br />

molecular biology,form the basis of a rapidly increasing<br />

number of systems in which DNA molecules are used to<br />

perform logic <strong>and</strong> computational functions,both in vitro, [599]<br />

<strong>and</strong> in living cells. [600] Such networks may prove important in<br />

the control of more sophisticated DNA-based mechanical<br />

devices.<br />

10. Conclusions <strong>and</strong> Outlook<br />

Angew<strong>and</strong>te<br />

Chemie<br />

In this Review we have outlined the current state-of-theart<br />

with regards to how the relative positioning of the<br />

components of molecular-level structures can be switched,<br />

rotated,speeded up,slowed down,<strong>and</strong> directionally driven in<br />

response to stimuli. In doing so they can affect the nanoscopic<br />

<strong>and</strong> macroscopic properties of the system to which they<br />

belong. Whether one chooses to call such structures “motors”<br />

<strong>and</strong> “machines”,or prefers to consider them more classically<br />

as specific triggered large amplitude conformational,configurational,<strong>and</strong><br />

structural changes,is irrelevant. What is<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

163


Reviews<br />

important is that the use of controlled molecular-level motion<br />

to bring about property changes is fundamentally different—<br />

both in terms of how to do it <strong>and</strong> in what it can do—to<br />

methods that rely solely upon variations in the nature of<br />

functional groups or electrostatics.<br />

Currently it is fair to say that there are two distinct design<br />

philosophies being used to try <strong>and</strong> prepare synthetic molecular-level<br />

machines: The first “hard matter” approach focuses<br />

on adapting mechanical principles <strong>and</strong> designs from the<br />

macroscopic world to molecules <strong>and</strong> is typified by the<br />

research efforts on field-driven rotors,molecular gyroscopes,<br />

molecular scissors,molecular wheelbarrows,nanocars,etc. In<br />

this approach,friction <strong>and</strong> binding interactions are avoided as<br />

much as possible <strong>and</strong> the molecules are designed to be rigid in<br />

all degrees of freedom except those involved in the desired<br />

motion. The input energy is designed to provide a directed<br />

force that pushes the molecular machine in the desired<br />

direction or exerts a torque to cause it to turn. This approach<br />

follows directly from that used in the design of macroscopic<br />

machines,which are designed to reduce friction <strong>and</strong> sticking<br />

<strong>and</strong> to eliminate jitter,vibration,<strong>and</strong> excess motion that<br />

dissipate energy <strong>and</strong> reduce the efficiency of the machine.<br />

The second “soft matter” approach is more chemical in<br />

philosophy <strong>and</strong> seeks to adapt principles of chemistry<br />

(selective stabilization <strong>and</strong> destabilization of noncovalent<br />

bonding interactions,structural symmetry,<strong>and</strong> kinetic versus<br />

thermodynamic control of processes) to achieve efficient <strong>and</strong><br />

controllable directional motion at the molecular level. This<br />

approach characterizes our own studies on catenanes <strong>and</strong><br />

rotaxanes,for example,<strong>and</strong> also that of the DNA walkers,etc.<br />

Directed motion is achieved by a sequence of reactions in<br />

which different covalent <strong>and</strong> noncovalent-bonding arrangements<br />

are alternately stabilized <strong>and</strong> destabilized. The input<br />

energy is used to ensure that the sequence of reactions occurs<br />

in one order <strong>and</strong> not the reverse. The motion occurs by<br />

diffusion over an energy barrier,where the conformational<br />

changes are best described as thermally activated transitions<br />

from states that are in local equilibrium.<br />

The advantage of the “hard matter” approach is that we<br />

underst<strong>and</strong> macroscopic mechanics very well <strong>and</strong> so can<br />

easily see how to construct mechanical machines. The<br />

disadvantage is that,under most conditions,such machines<br />

must fight against the physics <strong>and</strong> chemistry intrinsic to their<br />

length scale. The advantage of the “soft matter” approach is<br />

that by using the principles outlined in Sections 1.4.2,1.4.3,<br />

<strong>and</strong> 4.4 we can see how to exploit the natural features of the<br />

nanoscale environment. Its disadvantage is that there is no<br />

familiar macroscopic model to follow <strong>and</strong> nature s exquisite<br />

working examples are too complex in their detail to provide<br />

us with more than broad clues at present.<br />

As outlined in this Review,both sets of design philosophies<br />

have already had many notable successes <strong>and</strong> they are<br />

not mutually exclusive. No doubt their combination will<br />

become increasingly important in the future.<br />

Unfortunately,the hype surrounding science fictional<br />

“surgical nanobots” <strong>and</strong> the hysteria of “grey goo”,has led to<br />

controversy <strong>and</strong> the questioning of expectations for synthetic<br />

molecular machines. In a commentary [601] on the exciting<br />

recent developments in catalytically driven microparticles<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

(Section 6.1) the question was posed “What is the problem<br />

that requires having a nanomotor?”. Perhaps the best way to<br />

appreciate the technological potential of controlled molecular-level<br />

motion it is to recognize that this question has been<br />

asked <strong>and</strong> answered repeatedly over four billion years of<br />

evolution with the result that nanomotors <strong>and</strong> molecular-level<br />

machines lie at the heart of virtually every significant<br />

biological process. Nature has not repeatedly chosen this<br />

solution without good reason. In stark contrast to biology,<br />

none of mankind s fantastic myriad of present day technologies<br />

(with the exception of liquid crystals) exploit controlled<br />

molecular-level motion in any way at all. When we learn how<br />

to build synthetic structures that can rectify r<strong>and</strong>om dynamic<br />

processes <strong>and</strong> interface their effects directly with other<br />

molecular-level substructures <strong>and</strong> the outside world,it has<br />

the potential to revolutionize every aspect of functional<br />

molecule <strong>and</strong> materials design. An improved underst<strong>and</strong>ing<br />

of physics <strong>and</strong> biology will surely also follow.<br />

In this regard,we do not subscribe to the view that it will<br />

be impossible for synthetic chemists to develop molecularlevel<br />

machine systems that use controlled motion to perform<br />

functions similar to those of biological machines simply<br />

because of the complexity of the latter (just as the tremendous<br />

successes in homogeneous <strong>and</strong> heterogeneous catalysis have<br />

been achieved without the need for the de novo design of<br />

enzymes). The first examples of molecular-level mechanical<br />

switching being used to perform functions on surfaces <strong>and</strong> in<br />

solution have already been demonstrated. However,a broad<br />

new range of synthetic strategies <strong>and</strong> methodologies are<br />

needed to take this further. It is only in the last decade that<br />

unnatural product synthesis has progressed to a level where<br />

the architectures necessary for biasing conformational <strong>and</strong> coconformational<br />

motions can be constructed. However,this is<br />

not,in itself,enough to produce anything other than the most<br />

basic molecular-machine systems. Any machine more sophisticated<br />

than a switch must operate by manipulating nonequilibrium<br />

conformations <strong>and</strong>/or co-conformations of a<br />

substrate or itself. Although many methods for turning<br />

binding interactions “on” <strong>and</strong> “off” have been developed,<br />

relatively little is known about how to vary the kinetics of<br />

binding in response to external stimuli,<strong>and</strong> virtually nothing<br />

about how to do so in systems where the substrate is restricted<br />

from exchanging with the bulk.<br />

Sections 2.1.2,2.2,<strong>and</strong> 4.6 discussed the first examples of<br />

unidirectional motors based on rotation around single,<br />

double,<strong>and</strong> mechanical bonds,respectively. The development<br />

of molecular-level systems which carry out functions by<br />

responding to stimuli through Boolean logic operations [252]<br />

has clear relevance to the future development of interfaced<br />

<strong>and</strong> compartmentalized molecular machines which are more<br />

sophisticated than the current generation of mechanical<br />

switches <strong>and</strong> motors. While current efforts in molecular<br />

logic have concentrated on supramolecular systems rather<br />

than mechanical motion,their success suggests a way in which<br />

relatively simple molecular-machine components might be<br />

connected to produce useful devices at the next level of<br />

complexity (Section 4.4).<br />

Developments in synthetic chemistry almost always go<br />

h<strong>and</strong>-in-h<strong>and</strong> with advances in measurement <strong>and</strong> instrumen-<br />

164 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

tation. Spectacular advances in spectroscopy [602] <strong>and</strong> singlemolecule<br />

manipulation techniques (see Section 7) have<br />

already had a major impact in the study of molecular-level<br />

machine systems <strong>and</strong> further breakthroughs appear more<br />

than likely over the next few years.<br />

Despite the intrinsic simplicity of the current generation<br />

of synthetic molecular machines,attention is already turning<br />

to their application to perform useful tasks. Inspiration <strong>and</strong><br />

assistance again comes from an increasing underst<strong>and</strong>ing of<br />

biological motor proteins <strong>and</strong> information processing,as well<br />

as the remarkable achievements of modern-day microelectronics<br />

<strong>and</strong> mechanical engineering. Yet we must bear in mind<br />

that mechanically based devices made by the synthetic<br />

chemist will have different characteristics,strengths,<strong>and</strong><br />

weaknesses to their counterparts at other length scales. As<br />

with many fundamental developments in science—from<br />

electricity to manned-flight,the computer <strong>and</strong> the internet—it<br />

is not clear at the very beginning (<strong>and</strong> make no<br />

mistake that we are,indeed,at the very beginning in terms of<br />

experimental systems) in exactly what ways synthetic molecular-level<br />

machines are going to change technology. In the<br />

short term,changing surface properties,switching,<strong>and</strong><br />

memory devices look particularly attractive. A greater<br />

challenge,however,is to transduce the energy input from<br />

an external source to a molecular machine into the directed<br />

transport of a cargo or the pumping of a substrate to reverse a<br />

chemical gradient. The organization <strong>and</strong> addressing of<br />

molecular machines on surfaces <strong>and</strong> at interfaces is an area<br />

where we can expect particularly important advances over the<br />

next few years.<br />

The breadth of this Review is testament to the multidisciplinary<br />

nature of the subject matter. We hope it will be of<br />

interest to biologists to see what insights simpler molecular<br />

analogues might have for uncovering the workings of complex<br />

motor proteins. We hope it will be of interest to mathematicians<br />

<strong>and</strong> physicists to see what synthetic chemists can make<br />

(<strong>and</strong> cannot yet make) <strong>and</strong> the molecular design strategies<br />

that can be gleaned from their theories. We hope it will be of<br />

interest to materials <strong>and</strong> surface scientists to see what<br />

molecular-machine structures are available,which they can<br />

consider trying to organize,characterize,<strong>and</strong> utilize. We hope<br />

it will be of interest to engineers to see the kinds of property<br />

effects possible,<strong>and</strong> the stimuli that can be used,with<br />

molecules that need to be interfaced with the outside world.<br />

But most of all we hope it will of interest to chemists—<br />

synthetic,organic,inorganic,physical,<strong>and</strong> theoretical—<strong>and</strong><br />

that it might inspire them to take up a challenge in which the<br />

rules of the game are only just being appreciated,<strong>and</strong> where<br />

there is much virgin territory to be trodden <strong>and</strong> major<br />

breakthroughs to be made.<br />

Addendum (September 22, 2006)<br />

In the time since this Review was accepted for publication,progress<br />

in the field has continued apace. The general<br />

importance <strong>and</strong> broad appeal of synthetic molecular<br />

machines has been reflected in the regular appearance of<br />

various short review articles focusing on the most recent<br />

Angew<strong>and</strong>te<br />

Chemie<br />

advances in specific areas of the field,including: conformational<br />

switching in cavit<strong>and</strong>s, [603] conformational control by<br />

stereochemical relays, [604] switching of chiral properties, [605]<br />

light-stimulated machines, [606] switchable molecular shuttles,<br />

[607] transition-metal-based molecular machines, [608] crystalline<br />

molecular machines, [609] self-locomotion, [610] electromechanical<br />

molecular electronic devices, [611] liquid-crystalline<br />

phase alignment, [612] liquid-crystalline elastomer actuators, [613]<br />

<strong>and</strong> biomotors used to power self-assembly processes. [614]<br />

Recent experiment <strong>and</strong> theory has shown that interesting<br />

effects arise for ratchet systems in which there are several<br />

interacting particles moving over the same potential-energy<br />

surface. [615] This brings to mind the almost complete absence<br />

of polyrotaxane architectures from current synthetic molecular<br />

machine designs <strong>and</strong> suggests that these interlocked<br />

architectures may provide interesting systems in the future.<br />

Even today,the long-studied concept of correlated conformational<br />

motions continues to generate significant interest <strong>and</strong><br />

advances, [616]<br />

including an investigation of a molecularorbital-controlled<br />

gearing effect during a sigmatropic rearrangement.<br />

[616c] The more recent concept of molecular “gyroscope”<br />

structures has also seen new examples reported,with<br />

potentially interesting dynamic features. [617] Observed at the<br />

single-molecule level using STM,rotation of a porphyrinbased<br />

molecule on a silver surface has been shown to be<br />

reversibly switched on by addition of a small “hub” molecule.<br />

[618] Using an overcrowded alkene,a three-state luminescent<br />

switch has been demonstrated,whereby red-,blue-,<strong>and</strong><br />

nonfluorescent states can be accessed by appropriate application<br />

of light,heat,<strong>and</strong> electrons. [619] Continuing the systematic<br />

investigation of unidirectional motors constructed from<br />

overcrowded alkenes,new members of the “second generation”<br />

series have been reported, [620] including the fastestrotating<br />

example yet.<br />

Underst<strong>and</strong>ing <strong>and</strong> characterization of isolated molecular<br />

mechanical switching units is now such that devices of greater<br />

complexity,incorporating a number of fundamental features,<br />

are within reach. Aida <strong>and</strong> co-workers have harnessed an<br />

azobenzene isomerization-induced change in conformation<br />

around a metallocene pivot to control the relative arrangement<br />

of two porphyrin substituents. These,in turn,can<br />

regulate the conformation of a kinetically bound guest. [621] In<br />

a related system,a photoresponsive guest has been used to<br />

switch the host molecule between two,kinetically stable,<br />

conformations. [622] An example of macroscopic-machineinspired<br />

design can be found in a report on the construction<br />

<strong>and</strong> solution-phase characterization of an ambitious new<br />

member in the molecular “nanocar” series that combines the<br />

design features of the original STM-tip-driven molecules with<br />

an overcrowded alkene molecular motor as a proposed<br />

propulsion mechanism. [623]<br />

In terms of rotaxane-based switches,recent advances<br />

include: the first quantitative force spectroscopy measurements<br />

on intercomponent interactions in a shuttle; [624] further<br />

characterization of amphiphilic shuttles arranged at the air–<br />

water interface [625] <strong>and</strong> the fabrication of solid-state electronic<br />

devices from monolayers; [626] theoretical <strong>and</strong> experimental<br />

work aimed at further clarifying the mechanism of operation<br />

for solid-state MSTJs; [627] <strong>and</strong> progress towards an electrically<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

165


Reviews<br />

powered actuator constructed from such rotaxanes. [628] In a<br />

truly novel addition to the dynamic interlocked molecules<br />

field,a rotaxane has been reported in which raising the<br />

temperature leads to reduced intercomponent motions,as a<br />

result of the reversible formation of imine bonds between the<br />

macrocycle <strong>and</strong> thread. [629]<br />

<strong>Mechanical</strong> [630]<br />

<strong>and</strong> electrostatic [631]<br />

telescoping of<br />

MWNTs has recently been exploited to create a tunable<br />

nanoresonator <strong>and</strong> a nanoelectromechanical switch,respectively.<br />

Polymer-based actuator materials have continued their<br />

march towards useful devices,making progress on several<br />

fronts. The development of stimuli-responsive hydrogels that<br />

are biochemically degradable [632] <strong>and</strong> that respond to important<br />

biological analytes in relevant media [633] should prove to<br />

be important advances. In terms of engineering existing<br />

actuator materials,highlights include the use of stimuliresponsive<br />

hydrogels to control the focal length of liquid<br />

microlenses [634] <strong>and</strong> the direct conversion of chemical potential<br />

into work using charge- <strong>and</strong> heat-powered actuator<br />

materials by creating fuel-cell electrodes from carbon nanotube<br />

sheets <strong>and</strong> a shape memory alloy,respectively. [635]<br />

In an interesting experiment,STM has been applied to<br />

investigate conformational changes in single chlorophyll-a<br />

molecules on injection of electrons,thus revealing a four-step<br />

switching scheme. [636] Further development of modified MscL<br />

channel proteins has led to a tunable pH-sensitive valve that<br />

may find application in drug-delivery systems, [637] while an<br />

entirely synthetic peptide has been used to form switchable<br />

channels which open in the presence of Fe(III) ions. [638] It has<br />

been shown that microtubules being transported down<br />

kinesin-coated channels can be directed at channel junctions<br />

by application of an electric field. [639] A pH-responsive<br />

conformational switch constructed from DNA has recently<br />

been attached to a solid substrate <strong>and</strong> coupled to an<br />

oscillating chemical reaction to power its operation. [640]<br />

Finally,a remarkable self-assembling system of aromatic<br />

chromophores has been shown to span a lipid bilayer <strong>and</strong><br />

generate a photoinduced electric potential. This was used to<br />

power the reduction of quinones held inside vesicles formed<br />

by the membrane,thereby resulting in a transmembrane<br />

proton gradient. Addition of an aromatic intercalator irreversibly<br />

disrupted the complex,turning it into an ion channel<br />

through which the proton gradient could equilibrate. [641] If this<br />

is not quite a synthetic ion pump,it is well on the way there!<br />

The first four decades of supramolecular chemistry were<br />

primarily concerned with the manipulation of molecular-level<br />

structures under thermodynamic control. [169] However,the<br />

most recent developments in molecular-level machines demonstrate<br />

yet again that to build sophisticated functional<br />

systems it will be necessary to develop a whole new field of<br />

supramolecular <strong>and</strong> molecular chemistry that operates,with<br />

control,far from equilibrium. The underlying principles that<br />

will allow chemists to do this are outlined in Sections 1.4.2,<br />

1.4.3,<strong>and</strong> 4.4.<br />

D.A.L. <strong>and</strong> F.Z. would like to thank all the members, past <strong>and</strong><br />

present, of their research groups for their ideas, imagination,<br />

<strong>and</strong> enthusiasmover the past 15 years. We also thank the UK<br />

<strong>and</strong> Italian funding bodies, the EU, <strong>and</strong> the Carnegie Trust for<br />

their generous funding of our research programs in the area of<br />

molecular motors <strong>and</strong> machines.<br />

Received: December 5,2005<br />

Published online: November 29,2006<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

[1] a) R. P. Feynman, Eng. Sci. 1960, 23,22 – 36; b) A. P. Davis,<br />

Nature 1999, 401,120 – 121; c) V. Balzani,A. Credi,F. M.<br />

Raymo,J. F. Stoddart,Angew. Chem. 2000, 112,3484 – 3530;<br />

Angew. Chem. Int. Ed. 2000, 39,3349 – 3391; d) B. L. Feringa,<br />

Nature 2000, 408,151 – 154; e) special issue on “<strong>Molecular</strong><br />

<strong>Machines</strong>”: Acc. Chem. Res. 2001, 34,409 – 522; f) volume on<br />

“<strong>Molecular</strong> <strong>Machines</strong> <strong>and</strong> <strong>Motors</strong>”: Struct. Bonding (Berlin)<br />

2001, 99,1 – 281; g) V. Balzani,A. Credi,M. Venturi,Chem.<br />

Eur. J. 2002, 8,5524 – 5532; h) V. Balzani,M. Venturi,A. Credi,<br />

<strong>Molecular</strong> Devices <strong>and</strong> <strong>Machines</strong>. A Journey into the Nanoworld,Wiley-VCH,Weinheim,2003;<br />

i) C. J. Easton,S. F.<br />

Lincoln,L. Barr,H. Onagi,Chem. Eur. J. 2004, 10,3120 –<br />

3128; j) C. P. M<strong>and</strong>l,B. König,Angew. Chem. 2004, 116,<br />

1650 – 1652; Angew. Chem. Int. Ed. 2004, 43,1622 – 1624;<br />

k) G. S. Kottas,L. I. Clarke,D. Horinek,J. Michl,Chem. Rev.<br />

2005, 105,1281 – 1376; l) K. Kinbara,T. Aida,Chem. Rev. 2005,<br />

105,1377 – 1400; m) E. R. Kay,D. A. Leigh in Functional<br />

Artificial Receptors (Eds.: T. Schrader,A. D. Hamilton),Wiley-<br />

VCH,Weinheim,2005,pp. 333 – 406; n) volume on “<strong>Molecular</strong><br />

<strong>Machines</strong>”: Top. Curr. Chem. 2005, 262,1 – 236; for a recent<br />

view of the molecular-machines field from the perspective of<br />

the physics community,see: o) special issue on “<strong>Molecular</strong><br />

<strong>Motors</strong>”: J. Phys.: Condens. Matter 2005, 17,S3661–S4024.<br />

[2] This is in no sense a criticism of the terminology used at the<br />

time to report these pioneering experiments,which filled a<br />

generation of chemists with excitement at the prospect of<br />

making molecular-sized machines. Language can only ever be<br />

used in the context of the knowledge of the day.<br />

[3] E. R. Kay,D. A. Leigh,Nature 2006, 440,286 – 287.<br />

[4] a) R. E. Smalley, Sci. Am. 2001, 285(3),68 – 69; b) G. M.<br />

Whitesides, Sci. Am. 2001, 285(3),70 – 75; c) point/counterpoint<br />

debate between R. E. Smalley <strong>and</strong> K. E. Drexler, Chem.<br />

Eng. News 2003, 81(48),37 – 42.<br />

[5] In fact,physicists sometimes differentiate thermal ratchets<br />

from motors in that the former do not bear a load. However,as<br />

they are mechanistically identical,this distinction is unnecessary<br />

for the design of chemical systems.<br />

[6] “Motor molecules” can also act as switches; the change in the<br />

distribution or position of the components can be used to affect<br />

a system as a function of state; see Sections 8.3.2 <strong>and</strong> 8.3.3.<br />

[7] a) R. Brown, Philos. Mag. 1828, 4,171 – 173; b) R. Brown,<br />

Edinb. New Philos. J. 1828, 5,358 – 371; c) R. Brown,Philos.<br />

Mag. 1829, 6,161 – 166; Brown s accounts of these investigations<br />

are also reprinted in d) R. Brown in The Miscellaneous<br />

Botanical Works of Robert Brown, Vol. 1 (Ed.: J. J. Bennett),<br />

Ray Society,London,1866,pp. 463 – 486; for a modern<br />

reconstruction of the original experiments,see: e) B. J. Ford,<br />

Microscope 1992, 40,235 – 241.<br />

[8] A. Einstein, Ann. Phys. 1905, 17,549 – 560; reprinted in<br />

Einstein s Annalen Papers: The Complete Collection 1901–<br />

1922 (Ed.: J. Renn),Wiley-VCH,Weinheim,2005,pp. 182 –<br />

193.<br />

[9] J. Perrin, Atoms (English Translation: D. L. Hammick),2nd<br />

English ed.,Constable & Co.,London,1923.<br />

[10] For more detailed accounts of these discoveries <strong>and</strong> their<br />

impact on the scientific world,see: a) M. D. Haw,J. Phys.:<br />

Condens. Matter 2002, 14,7769 – 7779; b) G. Parisi,Nature 2005,<br />

433,221 – 221; c) M. Haw,Phys. World 2005, 18(1),19 – 22; d) P.<br />

Hänggi,F. Marchesoni,Chaos 2005, 15,026101; e) J. Renn,<br />

Ann. Phys. 2005, 14,S23-S37; reprinted in Einstein s Annalen<br />

166 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Papers: The Complete Collection 1901–1922 (Ed.: J. Renn),<br />

Wiley-VCH,Weinheim,2005,pp. 23 – 37.<br />

[11] <strong>Molecular</strong> <strong>Motors</strong> (Ed.: M. Schliwa),Wiley-VCH,Weinheim,<br />

2003.<br />

[12] R. D. Astumian,P. Hänggi,Phys. Today 2002, 55(11),33 – 39.<br />

[13] R. A. L. Jones, Soft <strong>Machines</strong>: Nanotechnology <strong>and</strong> Life,<br />

Oxford University Press,Oxford,2004.<br />

[14] For reprints of key papers <strong>and</strong> commentary on some of the<br />

main issues regarding Maxwell s demon,see: Maxwell s Demon<br />

2. Entropy, Classical <strong>and</strong> QuantumInformation, Computing<br />

(Eds.: H. S. Leff,A. F. Rex),Institute of Physics Publishing,<br />

Bristol, 2003.<br />

[15] The first a) private <strong>and</strong> b) public written discussions of the<br />

“temperature demon” were: a) J. C. Maxwell, Letter to P. G.<br />

Tait, 11 December 1867; quoted in C. G. Knott, Life <strong>and</strong><br />

Scientific Work of Peter Guthrie Tait,Cambridge University<br />

Press,London,1911,pp. 213 – 214; <strong>and</strong> reproduced in The<br />

Scientific Letters <strong>and</strong> Papers of James Clerk Maxwell Vol. II<br />

1862–1873 (Ed.: P. M. Harman),Cambridge University Press,<br />

Cambridge, 1995,pp. 331 – 332; b) J. C. Maxwell,Theory of<br />

Heat,Longmans,Green <strong>and</strong> Co.,London,1871,chap. 22;<br />

c) Maxwell introduced the idea of a “pressure demon” in a later<br />

letter to Tait (believed to date from early 1875); quoted in C. G.<br />

Knott, Life <strong>and</strong> Scientific Work of Peter Guthrie Tait,Cambridge<br />

University Press,London,1911,pp. 214 – 215; <strong>and</strong><br />

reproduced in The Scientific Letters <strong>and</strong> Papers of James<br />

Clerk Maxwell, Vol. III, 1874–1879 (Ed.: P. M. Harman),<br />

Cambridge University Press,Cambridge,2002,pp. 185 – 187;<br />

“Concerning Demons… Is the production of an inequality of<br />

temperature their only occupation? No,for less intelligent<br />

demons can produce a difference in pressure as well as<br />

temperature by merely allowing all particles going in one<br />

direction while stopping all those going the other way. This<br />

reduces the demon to a valve.”. More formally,a pressure<br />

demon would operate in a system linked to a constanttemperature<br />

reservoir with the sole effect of using energy<br />

transferred as heat from that reservoir to do work (see Szilard s<br />

engine,Figure 3 <strong>and</strong> Ref. [17]). This is in conflict with the<br />

Kelvin–Planck form of the second law of thermodynamics,<br />

whereas the temperature demon challenges the Clausius<br />

definition.<br />

[16] a) M. von Smoluchowski, Physik. Z. 1912, 13,1069-1080; b) M.<br />

von Smoluchowski, Vortrage über die Kinetische Theorie der<br />

Materie und der Elektrizitat (Ed.: M. Planck),Teubner <strong>and</strong><br />

Leipzig,Berlin,1914,pp. 89-121; Maxwell had earlier alluded to<br />

the basic premise behind Smoluchowski s trapdoor in private<br />

correspondence (c) J. C. Maxwell, Letter to J. W. Strutt, 6<br />

December 1870. Reproduced in The Scientific Letters <strong>and</strong><br />

Papers of James Clerk Maxwell Vol. II 1862-1873 (Ed.: P. M.<br />

Harman),Cambridge University Press,Cambridge,1995,pp.<br />

582-583),describing his “finite being” as a “doorkeeper,very<br />

intelligent <strong>and</strong> exceedingly quick” adding that “I do not see<br />

why even intelligence might not be dispensed with <strong>and</strong> the thing<br />

be made self-acting”.<br />

[17] L. Szilard, Z. Phys. 1929, 53,840 – 856. An English translation<br />

by A. Rapoport <strong>and</strong> M. Knoller is reprinted in ref. [14],<br />

pp. 110–119.<br />

[18] R. P. Feynman,R. B. Leighton,M. S<strong>and</strong>s,The Feynman<br />

Lectures on Physics, Vol. 1,Addison-Wesley,Reading,MA,<br />

1963,chapt. 46.<br />

[19] The Second Law of Thermodynamics. Memoirs by Carnot,<br />

Clausius <strong>and</strong> Thomson: Harper s Scientific Memoirs (Ed.: W. F.<br />

Magie),Harper & Brothers,New York,1899.<br />

[20] a) W. Ehrenberg, Sci. Am. 1967, 217(5),103 – 110; b) C. H.<br />

Bennett, Sci. Am. 1987, 257(5),88 – 96.<br />

[21] W. Thomson, Proc. R. Soc. Edinburgh 1874, 8,325 – 334.<br />

[22] W. Thomson, Nature 1874, 9,441 – 444.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

[23] In fact,Thomson thought that Maxwell had coined the term<br />

“demon”,but Maxwell corrects this in Ref. [15c].<br />

[24] This insight is all the more remarkable coming as it did before<br />

the advent of the electronic computer. The deconstruction of<br />

the reasoning processes of a living being to simple computational<br />

steps is the basis for the modern field of cybernetics.<br />

[25] For a modern critique of Feynman s analysis,see: a) J. M. R.<br />

Parrondo,P. Espaæol,Am. J. Phys. 1996, 64,1125 – 1130; b) K.<br />

Sekimoto, J. Phys. Soc. Jpn. 1997, 66,1234 – 1237; <strong>and</strong> for other<br />

modern treatments of this model,see: c) H. Sakaguchi,J. Phys.<br />

Soc. Jpn. 1998, 67,709 – 712; d) T. Hondou,F. Takagi,J. Phys.<br />

Soc. Jpn. 1998, 67,2974 – 2976; e) M. O. Magnasco,G. Stolovitzky,<br />

J. Stat. Phys. 1998, 93,615 – 632; f) C. Jarzynski,O.<br />

Mazonka, Phys. Rev. E 1999, 59,6448 – 6459; g) D. Abbott,<br />

B. R. Davis,J. M. R. Parrondo,in Proceedings of the Second<br />

International Conference on Unsolved Problems of Noise <strong>and</strong><br />

Fluctuations (Adelaide,1999),AIP Conf. Proc. 511 (Eds.: D.<br />

Abbott,L. B. Kiss),American Institute of Physics,New York,<br />

2000,pp. 213 – 218; h) S. Velasco,J. M. M. Roco,A. Medina,<br />

A. C. Hern<strong>and</strong>ez, J. Phys. D 2001, 34,1000 – 1006; i) R. Kawai,<br />

X. Sailer,L. Schimansky-Geier,C. Van den Broeck,Phys. Rev.<br />

E 2004, 69,051109; j) C. Van den Broeck,R. Kawai,P. Meurs,<br />

Phys. Rev. Lett. 2004, 93,090601; k) C. Van den Broeck,P.<br />

Meurs,R. Kawai,New J. Phys. 2005, 7,10.<br />

[26] In more recent times,Skordos <strong>and</strong> Zurek have carried out a<br />

simulation of the Smoluchowski trapdoor which showed<br />

Smoluchowski s intuitive conclusions (<strong>and</strong> Feynman s theoretical<br />

considerations) to be correct: a) P. A. Skordos,W. H.<br />

Zurek, Am. J. Phys. 1992, 60,876 – 882; reprinted in Ref. [14],<br />

pp. 94–100; cooling the trapdoor every time it approaches the<br />

closed position results in the success of the ratcheting<br />

mechanism <strong>and</strong> creates a pressure gradient between the two<br />

compartments. Rex <strong>and</strong> Larsen have presented a similar<br />

analysis in which they quantified the statistical entropy loss<br />

on sorting the particles <strong>and</strong> showed this to be less than the<br />

thermodynamic entropy gained on cooling the trapdoor:<br />

b) A. F. Rex,R. Larsen in Proceedings of the Workshop on<br />

Physics <strong>and</strong> Computation,IEEE Computer Society Press,Los<br />

Alamitos (PhysComp,Dallas,1992),1992,pp. 93–101;<br />

reprinted in Ref. [14],pp. 101–109.<br />

[27] E. M. Purcell, Am. J. Phys. 1977, 45,3 – 11.<br />

[28] J. Siegel, Science 2005, 310,63 – 64.<br />

[29] R. B. Gennis, Biomembranes—<strong>Molecular</strong> Structure <strong>and</strong> Function,Springer,New<br />

York,1989.<br />

[30] a) P. Läuger, Electrogenic Ion Pumps,Sinauer Associates,<br />

Sunderl<strong>and</strong>,MA,1991; b) E. Gouaux,R. MacKinnon,Science<br />

2005, 310,1461 – 1465.<br />

[31] a) C. Toyoshima,G. Inesi,Annu. Rev. Biochem. 2004, 73,269 –<br />

292; b) J. V. Møller,P. Nissen,T. L.-M. Sørensen,M. le Maire,<br />

Curr. Opin. Struct. Biol. 2005, 15,387 – 393.<br />

[32] M. N. Chatterjee,E. R. Kay,D. A. Leigh,J. Am. Chem. Soc.<br />

2006, 128,4058 – 4073.<br />

[33] D. G. Nicholls,S. J. Ferguson,Bioenergetics 3,Academic Press,<br />

London, 2002.<br />

[34] The precise mechanisms in each case are,of course,complex<br />

<strong>and</strong> the details difficult to elucidate. In part this is because<br />

elements of a conformational <strong>and</strong> a gradient-driven pumping<br />

mechanism can be in operation at the same time.<br />

[35] A close analogue,halorhodopsin,is a light-powered conformational<br />

chloride pump.<br />

[36] a) J. K. Lanyi,H. Luecke,Curr. Opin. Struct. Biol. 2001, 11,<br />

415 – 419; b) R. Neutze,E. Pebay-Peyroula,K. Edman,A.<br />

Royant,J. Navarro,E. M. L<strong>and</strong>au,Biochim. Biophys. Acta<br />

2002, 1565,144 – 167.<br />

[37] L. Onsager, Phys. Rev. 1931, 37,405 – 426.<br />

[38] For an introduction to Brownian motors,see: Ref. [12] <strong>and</strong><br />

R. D. Astumian, Sci. Am. 2001, 285(1),56 – 64.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

167


Reviews<br />

[39] For reviews on Brownian ratchet mechanisms,see: a) P.<br />

Hänggi,R. Bartussek in Nonlinear Physics of Complex<br />

Systems—Current status <strong>and</strong> future trends: Lecture Notes in<br />

Physics, Vol. 476 (Eds.: J. Parisi,S. C. Müller,W. Zimmermann),Springer,Berlin,1996,pp.<br />

294 – 308; b) R. D. Astumian,<br />

Science 1997, 276,917 – 922; c) M. Bier,Contemp. Phys. 1997,<br />

38,371 – 379; d) F. Jülicher,A. Ajdari,J. Prost,Rev. Mod. Phys.<br />

1997, 69,1269 – 1281; e) special issue on “The constructive role<br />

of noise in fluctuation-driven transport <strong>and</strong> stochastic resonance”:<br />

Chaos 1998, 8,533 – 664; f) P. Reimann,Phys. Rep.<br />

2002, 361,57 – 265; g) P. Reimann,P. Hänggi,Appl. Phys. A<br />

2002, 75,169 – 178; h) J. M. R. Parrondo,B. J. De Cisneros,<br />

Appl. Phys. A 2002, 75,179 – 191; i) J. M. R. Parrondo,L. Dinís,<br />

Contemp. Phys. 2004, 45,147 – 157; j) B. J. Gabrys´,K. Pesz,S. J.<br />

Bartkiewicz, Physica A 2004, 336,112 – 122; k) H. Linke,M. T.<br />

Downton,M. J. Zuckermann,Chaos 2005, 15,026111. The<br />

terms “thermal ratchet” <strong>and</strong> “Brownian ratchet” are often used<br />

interchangeably in the physics <strong>and</strong> biophysics literature<br />

(“Brownian rectifier” <strong>and</strong> “stochastic ratchet” are also occasionally<br />

used). This reflects the fact that thermal energy,in the<br />

form of Brownian motion,is the source of r<strong>and</strong>omization in<br />

such schemes. Although the latter term is more common in the<br />

recent literature,it has to be said that some researchers (see,for<br />

example,Refs. [39f–h]) prefer the former on the basis that<br />

“Brownian ratchet” was first coined (l) S. M. Simon,C. S.<br />

Peskin,G. F. Oster,Proc. Natl. Acad. Sci. USA 1992, 89,3770-<br />

3774) for the specific phenomenon of protein translocation<br />

through pores,during which Brownian motion is ratcheted by<br />

binding events on one side of the membrane. There can also be<br />

confusion between the terms “thermal ratchet” <strong>and</strong> “temperature<br />

ratchet”; the latter should properly be reserved only for<br />

ratchet systems that operate on the basis of temperature<br />

gradients,a subclass of thermal ratchets. We believe,therefore,<br />

that both force of numbers <strong>and</strong> the illustrative nature of the<br />

term “Brownian ratchet” commend it for use here,particularly<br />

as “Brownian motor” is commonly used to describe a machine<br />

which operates by such a mechanism (see Ref. [5]).<br />

[40] R<strong>and</strong>omizing influences other than Brownian motion can be<br />

used in ratchet mechanisms; for example,in quantum ratchets,<br />

quantum effects such as tunneling or electron-wave interference<br />

play this role; see,for example: H. Linke,T. E. Humphrey,<br />

P. E. Lindelof,A. Lofgren,R. Newbury,P. Omling,A. O.<br />

Sushkov,R. P. Taylor,H. Xu,Appl. Phys. A 2002, 75,237 – 246.<br />

[41] The source of spatial inversion symmetry breaking is most<br />

commonly an inherently asymmetric (or “ratchet”) potential.<br />

In certain cases,however,a symmetric potential is sufficient if<br />

the energy input (although unbiased) brings about spatial<br />

asymmetry. Also possible is spontaneous symmetry breaking in<br />

coupled symmetric systems driven away from equilibrium (not<br />

discussed here). Of course,spatial inversion symmetry may also<br />

be broken by a combination of the above.<br />

[42] a) R. D. Astumian,M. Bier,Biophys. J. 1996, 70,637 – 653;<br />

b) R. D. Astumian,I. DerØnyi,Eur. Biophys. J. 1998, 27,474 –<br />

489; c) L. Mahadevan,P. Matsudaira,Science 2000, 288,95 – 99;<br />

d) R. Lipowsky in Stochastic Processes in Physics, Chemistry<br />

<strong>and</strong> Biology: Lecture Notes in Physics, Vol. 557 (Eds.: J. A.<br />

Freund,T. Pöschel),Springer,Berlin,2000,pp. 21 – 31; e) C.<br />

Bustamante,D. Keller,G. Oster,Acc. Chem. Res. 2001, 34,<br />

412 – 420; f) R. D. Astumian, Appl. Phys. A 2002, 75,193 – 206;<br />

g) A. Mogilner,G. Oster,Curr. Biol. 2003, 13,R721 – R733;<br />

h) G. Oster,H. Y. Wang,Trends Cell Biol. 2003, 13,114 – 121;<br />

i) M. Kurzyński,P. Chełminiak,Physica A 2004, 336,123 – 132.<br />

[43] For a recent snapshot of the current status of experiment <strong>and</strong><br />

application with respect to both biological systems <strong>and</strong> future<br />

technology,see: special issue on “Ratchets <strong>and</strong> Brownian<br />

motors: Basics,experiments <strong>and</strong> applications”: Appl. Phys. A<br />

2002, 75,167 – 352.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

[44] For a selection of experimental examples,see: a) J. Rousselet,<br />

L. Salome,A. Ajdari,J. Prost,Nature 1994, 370,446 – 448;<br />

b) L. P. Faucheux,L. S. Bourdieu,P. D. Kaplan,A. J. Libchaber,<br />

Phys. Rev. Lett. 1995, 74,1504 – 1507; c) L. P. Faucheux,G.<br />

Stolovitzky,A. Libchaber,Phys. Rev. E 1995, 51,5239 – 5250;<br />

d) L. P. Faucheux,A. Libchaber,J. Chem. Soc. Faraday Trans.<br />

1995,3163 – 3166; e) L. Gorre,E. Ioannidis,P. Silberzan,<br />

Europhys. Lett. 1996, 33,267 – 272; f) L. Gorre-Talini,S.<br />

Jeanjean,P. Silberzan,Phys. Rev. E 1997, 56,2025 – 2034;<br />

g) L. Gorre-Talini,P. Silberzan,J. Phys. I 1997, 7,1475 – 1485;<br />

h) L. Gorre-Talini,J. P. Spatz,P. Silberzan,Chaos 1998, 8,650 –<br />

656; i) A. Lorke,S. Wimmer,B. Jager,J. P. Kotthaus,W.<br />

Wegscheider,M. Bichler,Physica B 1998, 251,312 – 316; j) H.<br />

Linke,W. Sheng,A. Lofgren,H. Q. Xu,P. Omling,P. E.<br />

Lindelof, Europhys. Lett. 1998, 44,341 – 347; k) C. Mennerat-<br />

Robilliard,D. Lucas,S. Guibal,J. Tabosa,C. Jurczak,J. Y.<br />

Courtois,G. Grynberg,Phys. Rev. Lett. 1999, 82,851 – 854; l) A.<br />

van Oudenaarden,S. G. Boxer,Science 1999, 285,1046 – 1048;<br />

m) J. S. Bader,R. W. Hammond,S. A. Henck,M. W. Deem,<br />

G. A. McDermott,J. M. Bustillo,J. W. Simpson,G. T. Mulhern,<br />

J. M. Rothberg, Proc. Natl. Acad. Sci. USA 1999, 96,13165 –<br />

13169; n) C. F. Chou,O. Bakajin,S. W. P. Turner,T. A. J. Duke,<br />

S. S. Chan,E. C. Cox,H. G. Craighead,R. H. Austin,Proc.<br />

Natl. Acad. Sci. USA 1999, 96,13762 – 13765; o) M. Switkes,<br />

C. M. Marcus,K. Campman,A. C. Gossard,Science 1999, 283,<br />

1905 – 1908; p) H. Linke,T. E. Humphrey,A. Lofgren,A. O.<br />

Sushkov,R. Newbury,R. P. Taylor,P. Omling,Science 1999,<br />

286,2314 – 2317; q) R. W. Hammond,J. S. Bader,S. A. Henck,<br />

M. W. Deem,G. A. McDermott,J. M. Bustillo,J. M. Rothberg,<br />

Electrophoresis 2000, 21,74 – 80; r) H. Linke,W. D. Sheng,A.<br />

Svensson,A. Lofgren,L. Christensson,H. Q. Xu,P. Omling,<br />

P. E. Lindelof, Phys. Rev. B 2000, 61,15914 – 15926; s) E. M.<br />

Hohberger,A. Lorke,W. Wegscheider,M. Bichler,Appl. Phys.<br />

Lett. 2001, 78,2905 – 2907; t) C. Marquet,A. Buguin,L. Talini,<br />

P. Silberzan, Phys. Rev. Lett. 2002, 88,168301; u) S. Matthias,F.<br />

Müller, Nature 2003, 424,53 – 57; v) L. R. Huang,E. C. Cox,<br />

R. H. Austin,J. C. Sturm,Anal. Chem. 2003, 75,6963 – 6967.<br />

The implications of Brownian ratchet theory,however,are farreaching<br />

<strong>and</strong> elements of a wide range of noise-assisted<br />

physical phenomena may be modeled as ratchet processes,<br />

see Refs. [39f,43].<br />

[45] a) H. X. Zhou,Y.-D. Chen,Phys. Rev. Lett. 1996, 77,194 – 197;<br />

b) I. DerØnyi,M. Bier,R. D. Astumian,Phys. Rev. Lett. 1999,<br />

83,903 – 906; c) A. Parmeggiani,F. Jülicher,A. Ajdari,J. Prost,<br />

Phys. Rev. E 1999, 60,2127 – 2140.<br />

[46] A power stroke is not intrinsically different in mechanistic<br />

terms to these ratchet mechanisms. The distinction is solely<br />

whether the particle is mostly under the influence of a force<br />

(that is,moving down an energy gradient) or not during its<br />

diffusion.<br />

[47] A net flux of particles may also be achieved in a symmetric<br />

periodic potential if it is subject to a spatially periodic variation<br />

in temperature of the same periodicity as,but out of phase with,<br />

the potential. Such systems have their origin in the demonstration<br />

by L<strong>and</strong>auer that in a bistable potential well,a<br />

localized inhomogeneity in temperature can alter the relative<br />

stabilities of the two minima (a) R. L<strong>and</strong>auer, Phys. Rev. A<br />

1975, 12,636 – 638); Büttiker (b) M. Büttiker,Z. Phys. B 1987,<br />

68,161 – 167); <strong>and</strong> van Kampen (c) N. G. van Kampen,IBM J.<br />

Res. Dev. 1988, 32,107 – 111) independently modeled this effect<br />

<strong>and</strong> showed that if extended to a periodic potential (that is,<br />

repeating the bistable well in a linear fashion or allowing<br />

motion in a closed loop),directional particle transport would<br />

occur; shortly after,L<strong>and</strong>auer further extended this model<br />

(d) R. L<strong>and</strong>auer, J. Stat. Phys. 1988, 53,233 – 248) which is now<br />

often termed “L<strong>and</strong>auer s blowtorch” (see also: e) K. Sinha,F.<br />

Moss, J. Stat. Phys. 1989, 54,1411 – 1423; f) M. Bekele,S.<br />

168 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Rajesh,G. Ananthakrishna,N. Kumar,Phys. Rev. E 1999, 59,<br />

143 – 149). Although realization of such a scheme on a microscopic<br />

scale would be problematic (see Section 1.3),circulating<br />

currents in a closed loop were first observed in electronics by<br />

Thomas Johann Seebeck in 1821. An electric current flows in a<br />

circuit composed of two resistors which are held at different<br />

temperatures (the modern equivalent of such thermoelectric<br />

circuits is a thermogenerator,in which a semiconductor diode is<br />

held at a different temperature to the rest of a circuit). For this<br />

reason,particle transport over symmetric periodic potentials,<br />

driven by spatial variations in temperature,is also often known<br />

as a “Seebeck ratchet”. Clearly,such schemes also bear a close<br />

resemblance to Feynman s ratchet-<strong>and</strong>-pawl (Section 1.2.1)<br />

when T 1 ¼6 T 2 <strong>and</strong>,indeed,mathematical equivalence of the<br />

two models can be shown,see ref. [25].<br />

[48] R. L<strong>and</strong>auer, IBM J. Res. Dev. 1961, 5,183 – 191.<br />

[49] The switching between two different intrinsic periodic potentials<br />

(chemical reactions) according to the position of the<br />

particle is effectively a chemical manifestation of “Parrondo s<br />

paradox”,a discovery in game theory that two losing repetitive<br />

gambling games,played stochastically or periodically one after<br />

another,can become winning; see Ref. [38] <strong>and</strong> a) P. V. E.<br />

McClintock, Nature 1999, 401,23 – 24; b) G. P. Harmer,D.<br />

Abbott, Nature 1999, 402,864; c) G. P. Harmer,D. Abbott,Stat.<br />

Sci. 1999, 14,206 – 213.<br />

[50] The directional threading that occurs in cyclodextrin <strong>and</strong> other<br />

unsymmetrical macrocycle-based rotaxanes,polyrotaxanes,<br />

<strong>and</strong> pseudorotaxanes could be used in this respect,but only if<br />

coupled to an appropriate energy input (for example,olefin or<br />

azobenzene isomerization),see: Section 4.3.1 <strong>and</strong> a) R. Isnin,<br />

A. E. Kaifer, J. Am. Chem. Soc. 1991, 113,8188 – 8190; b) H.<br />

Yonemura,M. Kasahara,H. Saito,H. Nakamura,T. Matsuo,J.<br />

Phys. Chem. 1992, 96,5765 – 5770; c) A. Abou-Hamdan,P.<br />

Bugnon,C. Saudan,P. G. Lye,A. E. Merbach,J. Am. Chem.<br />

Soc. 2000, 122,592 – 602; d) J. E. H. Buston,J. R. Young,H. L.<br />

Anderson, Chem. Commun. 2000,905 – 906; e) M. R. Craig,<br />

M. G. Hutchings,T. D. W. Claridge,H. L. Anderson,Angew.<br />

Chem. 2001, 113,1105 – 1108; Angew. Chem. Int. Ed. 2001, 40,<br />

1071 – 1074; f) C. Saudan,F. A. Dun<strong>and</strong>,A. Abou-Hamdan,P.<br />

Bugnon,P. G. Lye,S. F. Lincoln,A. E. Merbach,J. Am. Chem.<br />

Soc. 2001, 123,10 290 – 10298; g) J. S. Lock,B. L. May,P.<br />

Clements,S. F. Lincoln,C. J. Easton,Org. Biomol. Chem. 2004,<br />

2,337 – 344; h) J. W. Park,H. J. Song,Org. Lett. 2004, 6,4869 –<br />

4872; i) J. W. Park,S. Y. Lee,H. J. Song,K. K. Park,J. Org.<br />

Chem. 2005, 70,9505 – 9513; j) T. Oshikiri,Y. Takashima,H.<br />

Yamaguchi,A. Harada,J. Am. Chem. Soc. 2005, 127,12186 –<br />

12187; k) Q.-C. Wang,Q. Ma,D.-H. Qu,H. Tian,Chem. Eur. J.<br />

2006, 12,1088 – 1096; l) A. G. Cheetham,M. G. Hutchings,<br />

T. D. W. Claridge,H. L. Anderson,Angew. Chem. 2006, 118,<br />

1626 – 1629; Angew. Chem. Int. Ed. 2006, 45,1596 – 1599; m) G.<br />

Wenz,B.-H. Han,A. Müller,Chem. Rev. 2006, 106,782 – 817.<br />

[51] J. V. Hernµndez,E. R. Kay,D. A. Leigh,Science 2004, 306,<br />

1532 – 1537.<br />

[52] It seems to us that these ideas could prove useful in underst<strong>and</strong>ing<br />

protein folding <strong>and</strong>,perhaps,how some chaperones,<br />

foldases,<strong>and</strong> prions function.<br />

[53] a) K. Mislow, Acc. Chem. Res. 1976, 9,26 – 33; b) Z. Rappoport,<br />

S. E. Biali, Acc. Chem. Res. 1988, 21,442 – 449; c) K. Mislow,<br />

Chemtracts: Org. Chem. 1989, 2,151 – 174; d) Z. Rappoport,<br />

S. E. Biali, Acc. Chem. Res. 1997, 30,307 – 314.<br />

[54] M. Ōki, Angew. Chem. 1976, 88,67 – 74; Angew. Chem. Int. Ed.<br />

Engl. 1976, 15,87 – 93.<br />

[55] For a joint summary of the work of the two research groups,<br />

see: H. Iwamura,K. Mislow,Acc. Chem. Res. 1988, 21,175 –<br />

182.<br />

[56] For highlights of the work by Iwamura <strong>and</strong> co-workers,see:<br />

a) Y. Kawada,H. Iwamura,J. Org. Chem. 1980, 45,2547 – 2548;<br />

Angew<strong>and</strong>te<br />

Chemie<br />

b) Y. Kawada,H. Iwamura,J. Am. Chem. Soc. 1981, 103,958 –<br />

960; c) Y. Kawada,H. Iwamura,Tetrahedron Lett. 1981, 22,<br />

1533 – 1536; d) Y. Kawada,H. Iwamura,J. Am. Chem. Soc.<br />

1983, 105,1449 – 1459; e) Y. Kawada,Y. Okamoto,H. Iwamura,<br />

Tetrahedron Lett. 1983, 24,5359 – 5362; f) H. Iwamura,T. Ito,<br />

H. Ito,K. Toriumi,Y. Kawada,E. Osawa,T. Fujiyoshi,C. Jaime,<br />

J. Am. Chem. Soc. 1984, 106,4712 – 4717; g) H. Iwamura,J.<br />

Mol. Struct. 1985, 126,401 – 412; h) N. Koga,H. Iwamura,J.<br />

Am. Chem. Soc. 1985, 107,1426 – 1427; i) N. Koga,H. Iwamura,<br />

Chem. Lett. 1986,247 – 250; j) Y. Kawada,H. Yamazaki,G.<br />

Koga,S. Murata,H. Iwamura,J. Org. Chem. 1986, 51,1472 –<br />

1477; k) Y. Kawada,J. Ishikawa,H. Yamazaki,G. Koga,S.<br />

Murata,H. Iwamura,Tetrahedron Lett. 1987, 28,445 – 448.<br />

[57] For highlights of the work by Mislow <strong>and</strong> co-workers,see:<br />

a) W. D. Hounshell,C. A. Johnson,A. Guenzi,F. Cozzi,K.<br />

Mislow, Proc. Natl. Acad. Sci. USA 1980, 77,6961 – 6964; b) F.<br />

Cozzi,A. Guenzi,C. A. Johnson,K. Mislow,W. D. Hounshell,<br />

J. F. Blount, J. Am. Chem. Soc. 1981, 103,957 – 958; c) C. A.<br />

Johnson,A. Guenzi,K. Mislow,J. Am. Chem. Soc. 1981, 103,<br />

6240 – 6242; d) C. A. Johnson,A. Guenzi,R. B. Nachbar,J. F.<br />

Blount,O. Wennerstrom,K. Mislow,J. Am. Chem. Soc. 1982,<br />

104,5163 – 5168; e) H. B. Burgi,W. D. Hounshell,R. B. Nachbar,K.<br />

Mislow,J. Am. Chem. Soc. 1983, 105,1427 – 1438; f) A.<br />

Guenzi,C. A. Johnson,F. Cozzi,K. Mislow,J. Am. Chem. Soc.<br />

1983, 105,1438 – 1448.<br />

[58] The term “gearing” or “static gearing” is sometimes used to<br />

describe systems in which steric crowding results in the twisting<br />

<strong>and</strong> intermeshing of substituents on a planar framework,so as<br />

to all twist in the same direction. This is not,however,a<br />

sufficient condition for dynamic gearing (that is,correlated<br />

motion) of these substituents to occur.<br />

[59] Like atropisomers,residual stereoisomers arise when two forms<br />

can interconvert through bond rotations but only at high<br />

energies. The term residual stereoisomer is used when each<br />

isolable species is not in fact a single conformational possibility,<br />

but rather a collection of rapidly interconverting permutations.<br />

Where these forms are interconverting by correlated rotations,<br />

the residual isomers may also be termed phase isomers.<br />

[60] a) N. Koga,Y. Kawada,H. Iwamura,J. Am. Chem. Soc. 1983,<br />

105,5498 – 5499; b) N. Koga,Y. Kawada,H. Iwamura,Tetrahedron<br />

1986, 42,1679 – 1686.<br />

[61] Y. Kawada,H. Sakai,M. Oguri,G. Koga,Tetrahedron Lett.<br />

1994, 35,139 – 142.<br />

[62] a) G. Yamamoto,M. Ōki, Chem. Lett. 1979,1251 – 1254; b) G.<br />

Yamamoto,M. Ōki, Chem. Lett. 1979,1255 – 1258; c) G.<br />

Yamamoto,M. Ōki, Bull. Chem. Soc. Jpn. 1981, 54,473 – 480;<br />

d) G. Yamamoto,M. Ōki, Bull. Chem. Soc. Jpn. 1981, 54,481 –<br />

487; e) G. Yamamoto,M. Ōki, J. Org. Chem. 1983, 48,1233 –<br />

1236; f) G. Yamamoto, J. Mol. Struct. 1985, 126,413 – 420; g) G.<br />

Yamamoto,M. Ōki, Bull. Chem. Soc. Jpn. 1985, 58,1953 – 1961;<br />

h) G. Yamamoto,M. Ōki, Bull. Chem. Soc. Jpn. 1986, 59,3597 –<br />

3603; i) G. Yamamoto, Bull. Chem. Soc. Jpn. 1989, 62,4058 –<br />

4060; j) G. Yamamoto, Tetrahedron 1990, 46,2761 – 2772.<br />

[63] A. M. Stevens,C. J. Richards,Tetrahedron Lett. 1997, 38,7805 –<br />

7808.<br />

[64] For a review of investigations aimed at uncovering correlated<br />

motion in other crowded organometallic systems,see: S.<br />

Brydges,L. E. Harrington,M. J. McGlinchey,Coord. Chem.<br />

Rev. 2002, 233,75 – 105.<br />

[65] a) S. Hiraoka,K. Harano,T. Tanaka,M. Shiro,M. Shionoya,<br />

Angew. Chem. 2003, 115,5340 – 5343; Angew. Chem. Int. Ed.<br />

2003, 42,5182 – 5185; b) S. Hiraoka,M. Shiro,M. Shionoya,J.<br />

Am. Chem. Soc. 2004, 126,1214 – 1218; c) S. Hiraoka,K.<br />

Hirata,M. Shionoya,Angew. Chem. 2004, 116,3902 – 3906;<br />

Angew. Chem. Int. Ed. 2004, 43,3814 – 3818.<br />

[66] a) M. Kuttenberger,M. Frieser,M. Hofweber,A. Mannschreck,<br />

Tetrahedron: Asymmetry 1998, 9,3629 – 3645; b) J.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

169


Reviews<br />

Clayden,J. H. Pink,Angew. Chem. 1998, 110,2040 – 2043;<br />

Angew. Chem. Int. Ed. 1998, 37,1937 – 1939; c) E. R. Johnston,<br />

R. Fortt,J. C. Barborak,Magn. Reson. Chem. 2000, 38,932 –<br />

936; d) R. A. Bragg,J. Clayden,Org. Lett. 2000, 2,3351 – 3354;<br />

e) R. A. Bragg,J. Clayden,G. A. Morris,J. H. Pink,Chem. Eur.<br />

J. 2002, 8,1279 – 1289.<br />

[67] For accounts of the development of the Kelly systems,see:<br />

a) T. R. Kelly, Acc. Chem. Res. 2001, 34,514 – 522; b) T. R.<br />

Kelly,J. P. Sestelo,Struct. Bonding (Berlin) 2001, 99,19 – 53;<br />

c) J. P. Sestelo,T. R. Kelly,Appl. Phys. A 2002, 75,337 – 343.<br />

[68] T. R. Kelly,M. C. Bowyer,K. V. Bhaskar,D. Bebbington,A.<br />

Garcia,F. R. Lang,M. H. Kim,M. P. Jette,J. Am. Chem. Soc.<br />

1994, 116,3657 – 3658.<br />

[69] Metal-ion binding to a 2,2’-bipyridyl unit has also been used to<br />

reversibly switch the orientation of two appended porphyrin<br />

units between a freely rotating <strong>and</strong> a rigid cofacial conformation,see:<br />

Y. Tomohiro,A. Satake,Y. Kobuke,J. Org. Chem.<br />

2001, 66,8442 – 8446.<br />

[70] P. V. Jog,R. E. Brown,D. K. Bates,J. Org. Chem. 2003, 68,<br />

8240 – 8243.<br />

[71] Rotation around C aryl N bonds can be controlled by electrochemical<br />

switching between p-phenylenediamine (benzoid<br />

forms,rotation allowed) <strong>and</strong> p-quinonediimine (quinoid<br />

forms,no rotation) functionalities. This has been used to<br />

reversibly alter the geometry of a cyclophane,see: H.<br />

Kanazawa,M. Higuchi,K. Yamamoto,J. Am. Chem. Soc.<br />

2005, 127,16 404 – 16405.<br />

[72] a) T. R. Kelly,I. Tellitu,J. P. Sestelo,Angew. Chem. 1997, 109,<br />

1969 – 1972; Angew. Chem. Int. Ed. Engl. 1997, 36,1866 – 1868;<br />

b) T. R. Kelly,J. P. Sestelo,I. Tellitu,J. Org. Chem. 1998, 63,<br />

3655 – 3665.<br />

[73] a) A. P. Davis, Angew. Chem. 1998, 110,953 – 954; Angew.<br />

Chem. Int. Ed. 1998, 37,909 – 910; b) G. Musser,Sci. Am. 1999,<br />

280(2),24.<br />

[74] P. W. Atkins, Physical Chemistry,6th ed.,Oxford University<br />

Press,Oxford,1998.<br />

[75] The experimentally observed behavior of molecule 10 can be<br />

considered as a verification of both the second law of<br />

thermodynamics <strong>and</strong> the principle of detailed balance (see<br />

Section 1.4.1); for analysis,see: K. L. Sebastian,Phys. Rev. E<br />

2000, 61,937 – 939.<br />

[76] a) T. R. Kelly,H. De Silva,R. A. Silva,Nature 1999, 401,150 –<br />

152; b) T. R. Kelly,R. A. Silva,H. De Silva,S. Jasmin,Y. J.<br />

Zhao, J. Am. Chem. Soc. 2000, 122,6935 – 6949.<br />

[77] W. L. Mock,K. J. Ochwat,J. Phys. Org. Chem. 2003, 16,175 –<br />

182.<br />

[78] The presence of a kinetically unrestricted intermediate (bottom<br />

center in Scheme 4) reduces the ability of any molecular motor<br />

to progressively do work. Although an overall bias in the<br />

direction of the half-rotation should result from the molecular<br />

chirality,any attempt to achieve this against an opposing force<br />

(that is,to do work) would result in equilibration at this freely<br />

rotating intermediate. In the current case,preactivation of the<br />

“free” carboxylate prior to hydrolysis of the anhydride reduces<br />

the chance of reversal to give the starting enantiomer (an idling<br />

step) while also increasing the rate of the desired ring-closing<br />

reaction,in an attempt to minimize the effect of thermal<br />

r<strong>and</strong>omization. However,it is likely that any opposing force<br />

would swamp this effect.<br />

[79] B. J. Dahl,B. P. Branchaud,Tetrahedron Lett. 2004, 45,9599 –<br />

9602.<br />

[80] Y. Lin,B. J. Dahl,B. P. Branchaud,Tetrahedron Lett. 2005, 46,<br />

8359 – 8362.<br />

[81] S. P. Fletcher,F. Dumur,M. M. Pollard,B. L. Feringa,Science<br />

2005, 310,80 – 82.<br />

[82] a) K. Tashiro,K. Konishi,T. Aida,Angew. Chem. 1997, 109,<br />

882 – 884; Angew. Chem. Int. Ed. Engl. 1997, 36,856 – 858; b) K.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Tashiro,T. Fujiwara,K. Konishi,T. Aida,Chem. Commun.<br />

1998,1121 – 1122; c) K. Tashiro,K. Konishi,T. Aida,J. Am.<br />

Chem. Soc. 2000, 122,7921 – 7926.<br />

[83] Conversely,complexes formed around lanthanum(III) have<br />

been shown to rotate faster than the cerium(IV) parent<br />

compounds,see: M. Ikeda,M. Takeuchi,S. Shinkai,F. Tani,<br />

Y. Naruta, Bull. Chem. Soc. Jpn. 2001, 74,739 – 746.<br />

[84] a) S. Shinkai,M. Ikeda,A. Sugasaki,M. Takeuchi,Acc. Chem.<br />

Res. 2001, 34,494 – 503; b) M. Takeuchi,M. Ikeda,A. Sugasaki,<br />

S. Shinkai, Acc. Chem. Res. 2001, 34,865 – 873; c) S. Shinkai,M.<br />

Takeuchi, Bull. Chem. Soc. Jpn. 2005, 78,40 – 51.<br />

[85] a) M. Takeuchi,T. Imada,S. Shinkai,Angew. Chem. 1998, 110,<br />

2242 – 2246; Angew. Chem. Int. Ed. 1998, 37,2096 – 2099; b) A.<br />

Sugasaki,M. Ikeda,M. Takeuchi,A. Robertson,S. Shinkai,J.<br />

Chem. Soc. Perkin Trans. 1 1999,3259 – 3264.<br />

[86] M. Yamamoto,A. Sugasaki,M. Ikeda,M. Takeuchi,K. Frimat,<br />

T. D. James,S. Shinkai,Chem. Lett. 2001,520 – 521.<br />

[87] A. Robertson,M. Ikeda,M. Takeuchi,S. Shinkai,Bull. Chem.<br />

Soc. Jpn. 2001, 74,883 – 888.<br />

[88] a) M. Ikeda,T. Tanida,M. Takeuchi,S. Shinkai,Org. Lett. 2000,<br />

2,1803 – 1805; b) M. Ikeda,M. Takeuchi,S. Shinkai,F. Tani,Y.<br />

Naruta,S. Sakamoto,K. Yamaguchi,Chem. Eur. J. 2002, 8,<br />

5542 – 5550.<br />

[89] a) A. Sugasaki,M. Ikeda,M. Takeuchi,K. Koumoto,S. Shinkai,<br />

Tetrahedron 2000, 56,4717 – 4723; b) A. Sugasaki,M. Ikeda,M.<br />

Takeuchi,S. Shinkai,Angew. Chem. 2000, 112,3997 – 4000;<br />

Angew. Chem. Int. Ed. 2000, 39,3839 – 3842; c) A. Sugasaki,K.<br />

Sugiyasu,M. Ikeda,M. Takeuchi,S. Shinkai,J. Am. Chem. Soc.<br />

2001, 123,10 239 – 10244.<br />

[90] G. Ercolani, Org. Lett. 2005, 7,803 – 805.<br />

[91] Allosteric receptors based on the braking of rotational motion<br />

have also been created from porphyrin-based systems,connected<br />

not by a common metal ion,but by single <strong>and</strong> triple C C<br />

bonds,see: a) M. Ikeda,S. Shinkai,A. Osuka,Chem. Commun.<br />

2000,1047 – 1048; b) Y. Kubo,M. Ikeda,A. Sugasaki,M.<br />

Takeuchi,S. Shinkai,Tetrahedron Lett. 2001, 42,7435 – 7438;<br />

c) Y. Kubo,A. Sugasaki,M. Ikeda,K. Sugiyasu,K. Sonoda,A.<br />

Ikeda,M. Takeuchi,S. Shinkai,Org. Lett. 2002, 4,925 – 928;<br />

d) M. Ayabe,A. Ikeda,Y. Kubo,M. Takeuchi,S. Shinkai,<br />

Angew. Chem. 2002, 114,2914 – 2916; Angew. Chem. Int. Ed.<br />

2002, 41,2790 – 2792.<br />

[92] R. K. Bohn,A. Haal<strong>and</strong>,J. Organomet. Chem. 1966, 5,470 –<br />

476.<br />

[93] X.-B. Wang,B. Dai,H.-K. Woo,L.-S. Wang,Angew. Chem.<br />

2005, 117,6176 – 6178; Angew. Chem. Int. Ed. 2005, 44,6022 –<br />

6024.<br />

[94] a) L. J. Todd, Adv. Organomet. Chem. 1970, 8,87 – 115; b) M. F.<br />

Hawthorne,G. B. Dunks,Science 1972, 178,462 – 471.<br />

[95] a) L. F. Warren,M. F. Hawthorne,J. Am. Chem. Soc. 1970, 92,<br />

1157 – 1173; b) D. St. Clair,A. Zalkin,D. H. Templeton,J. Am.<br />

Chem. Soc. 1970, 92,1173 – 1179; c) M. R. Churchill,K. Gold,J.<br />

Am. Chem. Soc. 1970, 92,1180 – 1187.<br />

[96] M. F. Hawthorne,J. I. Zink,J. M. Skelton,M. J. Bayer,C. Liu,<br />

E. Livshits,R. Baer,D. Neuhauser,Science 2004, 303,1849 –<br />

1851.<br />

[97] For reviews on triggering protein folding events to allow the<br />

study of this otherwise natural phenomenon,see: a) W. A.<br />

Eaton,P. A. Thompson,C. K. Chan,S. J. Hagen,J. Hofrichter,<br />

Structure 1996, 4,1133 – 1139; b) special issue on “Protein<br />

Folding”: Acc. Chem. Res. 1998, 31,697 – 780; c) R. H. Callender,R.<br />

B. Dyer,R. Gilmanshin,W. H. Woodruff,Annu. Rev.<br />

Phys. Chem. 1998, 49,173 – 202; d) C. M. Dobson,A. S ˇ ali,M.<br />

Karplus, Angew. Chem. 1998, 110,908 – 935; Angew. Chem. Int.<br />

Ed. 1998, 37,868 – 893; e) M. Gruebele,Annu. Rev. Phys.<br />

Chem. 1999, 50,485 – 516; f) D. J. Brockwell,D. A. Smith,S. E.<br />

Radford, Curr. Opin. Struct. Biol. 2000, 10,16 – 25; g) Mechanisms<br />

of Protein Folding: Frontiers in <strong>Molecular</strong> Biology,<br />

170 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Vol. 32,2nd ed. (Ed.: R. H. Pain),Oxford University Press,<br />

Oxford, 2000; h) M. Volk, Eur. J. Org. Chem. 2001,2605 – 2621;<br />

i) A. R. Fersht,V. Daggett,Cell 2002, 108,573 – 582; for reviews<br />

on stimuli-induced control of synthetic polypeptide <strong>and</strong> polysaccharide<br />

secondary structures,see: j) M. Sisido,Prog. Polym.<br />

Sci. 1992, 17,699 – 764; k) D. W. Urry,Angew. Chem. 1993, 105,<br />

859 – 883; Angew. Chem. Int. Ed. Engl. 1993, 32,819 – 841;<br />

l) D. M. Whitfield,S. Stojkovski,B. Sarkar,Coord. Chem. Rev.<br />

1993, 122,171 – 225; m) J. P. Schneider,J. W. Kelly,Chem. Rev.<br />

1995, 95,2169 – 2187; n) T. Kinoshita,Prog. Polym. Sci. 1995,<br />

20,527 – 583; o) I. Willner,S. Rubin,Angew. Chem. 1996, 108,<br />

419 – 439; Angew. Chem. Int. Ed. Engl. 1996, 35,367 – 385; p) I.<br />

Willner, Acc. Chem. Res. 1997, 30,347 – 356; q) O. Pieroni,A.<br />

Fissi,G. Popova,Prog. Polym. Sci. 1998, 23,81 – 123; r) O.<br />

Pieroni,A. Fissi,N. Angelini,F. Lenci,Acc. Chem. Res. 2001,<br />

34,9 – 17; s) I. Willner,B. Willner in <strong>Molecular</strong> Switches (Ed.:<br />

B. L. Feringa),Wiley-VCH,Weinheim,2001,pp. 165 – 218; t) F.<br />

Ciardelli,O. Pieroni in <strong>Molecular</strong> Switches (Ed.: B. L. Feringa),<br />

Wiley-VCH,Weinheim,2001,pp. 399 – 441; u) G. A. Woolley,<br />

Acc. Chem. Res. 2005, 38,486 – 493.<br />

[98] a) A. Fersht, Enzyme Structure <strong>and</strong> Mechanism,2nd ed.,W. H.<br />

Freeman <strong>and</strong> Company,New York,1985; b) M. F. Perutz, Q.<br />

Rev. Biophys. 1989, 22,139 – 236; c) A. Mattevi,M. Rizzi,M.<br />

Bolognesi, Curr. Opin. Struct. Biol. 1996, 6,824 – 829; d) J.-P.<br />

Changeux,S. J. Edelstein,Neuron 1998, 21,959 – 980; e) W. A.<br />

Lim, Curr. Opin. Struct. Biol. 2002, 12,61 – 68; f) J. A. Hardy,<br />

J. A. Wells, Curr. Opin. Struct. Biol. 2004, 14,706 – 715; g) J.-P.<br />

Changeux,S. J. Edelstein,Science 2005, 308,1424 – 1428.<br />

[99] a) J. Rebek,Jr., Acc. Chem. Res. 1984, 17,258 – 264; b) F.<br />

Vanveggel,W. Verboom,D. N. Reinhoudt,Chem. Rev. 1994, 94,<br />

279 – 299; c) B. Linton,A. D. Hamilton,Chem. Rev. 1997, 97,<br />

1669 – 1680; d) A. Robertson,S. Shinkai,Coord. Chem. Rev.<br />

2000, 205,157 – 199; e) S. Shinkai in <strong>Molecular</strong> Switches (Ed.:<br />

B. L. Feringa),Wiley-VCH,Weinheim,2001,pp. 281 – 307; f) L.<br />

Kovbasyuk,R. Krämer,Chem. Rev. 2004, 104,3161 – 3187.<br />

[100] a) J. Rebek,Jr., J. E. Trend, R. V. Wattley, S. Chakravorti, J.<br />

Am. Chem. Soc. 1979, 101,4333 – 4337; b) J. Rebek,Jr.,R. V.<br />

Wattley, J. Am. Chem. Soc. 1980, 102,4853 – 4854.<br />

[101] J. Rebek,Jr.,T. Costello,L. Marshall,R. Wattley,R. C.<br />

Gadwood,K. Onan,J. Am. Chem. Soc. 1985, 107,7481 – 7487.<br />

[102] a) F.-G. Klärner,B. Kahlert,Acc. Chem. Res. 2003, 36,919 –<br />

932; b) M. Harmata, Acc. Chem. Res. 2004, 37,862 – 873.<br />

[103] a) R. P. Sijbesma,S. S. Wijmenga,R. J. M. Nolte,J. Am. Chem.<br />

Soc. 1992, 114,9807 – 9813; b) J. N. H. Reek,H. Engelkamp,<br />

A. E. Rowan,J. Elemans,R. J. M. Nolte,Chem. Eur. J. 1998, 4,<br />

716 – 722; c) A. E. Rowan,J. Elemans,R. J. M. Nolte,Acc.<br />

Chem. Res. 1999, 32,995 – 1006.<br />

[104] R. P. Sijbesma,R. J. M. Nolte,J. Am. Chem. Soc. 1991, 113,<br />

6695 – 6696.<br />

[105] a) Y. S. Chong,M. D. Smith,K. D. Shimizu,J. Am. Chem. Soc.<br />

2001, 123,7463 – 7464; b) C. F. Degenhardt,J. M. Lavin,M. D.<br />

Smith,K. D. Shimizu,Org. Lett. 2005, 7,4079 – 4081.<br />

[106] a) P. Leighton,J. K. M. S<strong>and</strong>ers,J. Chem. Soc. Chem. Commun.<br />

1984,854 – 856; b) R. J. Abraham,P. Leighton,J. K. M. S<strong>and</strong>ers,<br />

J. Am. Chem. Soc. 1985, 107,3472 – 3478.<br />

[107] E. L. Eliel,S. H. Wilen,Stereochemistry of Organic Compounds,Wiley-Interscience,New<br />

York,1994.<br />

[108] For reviews on conformational design,see: a) R. W. Hoffmann,<br />

Angew. Chem. 1992, 104,1147 – 1157; Angew. Chem. Int. Ed.<br />

Engl. 1992, 31,1124 – 1134; b) R. W. Hoffmann,Angew. Chem.<br />

2000, 112,2134 – 2150; Angew. Chem. Int. Ed. 2000, 39,2054 –<br />

2070.<br />

[109] a) H. Yuasa,H. Hashimoto,J. Am. Chem. Soc. 1999, 121,5089 –<br />

5090; b) T. Izumi,H. Hashimoto,H. Yuasa,Chem. Commun.<br />

2004,94 – 95.<br />

[110] a) F. M. Menger,P. A. Chicklo,M. J. Sherrod,Tetrahedron Lett.<br />

1989, 30,6943 – 6946; b) S. M. Shirodkar,G. R. Weisman,J.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Chem. Soc. Chem. Commun. 1989,236 – 238; c) M. Raban,<br />

D. L. Burch,E. R. Hortelano,D. Durocher,D. Kost,J. Org.<br />

Chem. 1994, 59,1283 – 1287; d) M. Goodall,P. M. Kelly,D.<br />

Parker,K. Gloe,H. Stephan,J. Chem. Soc. Perkin Trans. 2 1997,<br />

59 – 69; e) V. A. Palyulin,S. V. Emets,V. A. Chertkov,C.<br />

Kasper,H.-J. Schneider,Eur. J. Org. Chem. 1999,3479 – 3482.<br />

[111] a) D. S. Kemp,K. S. Petrakis,J. Org. Chem. 1981, 46,5140 –<br />

5143; b) V. V. Samoshin,V. A. Chertkov,L. P. Vatlina,E. K.<br />

Dobretsova,N. A. Simonov,L. P. Kastorsky,D. E. Gremyachinsky,H.-J.<br />

Schneider,Tetrahedron Lett. 1996, 37,3981 –<br />

3984; c) A. G. D. Santos,W. Klute,J. Torode,V. P. W. Bohm,<br />

E. Cabrita,J. Runsink,R. W. Hoffmann,New J. Chem. 1998, 22,<br />

993 – 997.<br />

[112] H. Yuasa,H. Hashimoto,Trends Glycosci. Glycotechnol. 2001,<br />

13,31 – 55.<br />

[113] For a review of resorcinarenes,including their conformational<br />

dynamics,see: P. Timmerman,W. Verboom,D. N. Reinhoudt,<br />

Tetrahedron 1996, 52,2663 – 2704.<br />

[114] a) J. R. Moran,S. Karbach,D. J. Cram,J. Am. Chem. Soc. 1982,<br />

104,5826 – 5828; b) J. R. Moran,J. L. Ericson,E. Dalcanale,<br />

J. A. Bryant,C. B. Knobler,D. J. Cram,J. Am. Chem. Soc. 1991,<br />

113,5707 – 5714.<br />

[115] a) P. J. Skinner,A. G. Cheetham,A. Beeby,V. Gramlich,F.<br />

Diederich, Helv. Chim. Acta 2001, 84,2146 – 2153; b) V. A.<br />

Azov,F. Diederich,Y. Lill,B. Hecht,Helv. Chim. Acta 2003, 86,<br />

2149 – 2155; c) V. A. Azov,B. Jaun,F. Diederich,Helv. Chim.<br />

Acta 2004, 87,449 – 462.<br />

[116] M. Frei,F. Marotti,F. Diederich,Chem. Commun. 2004,1362 –<br />

1363.<br />

[117] a) F. C. Tucci,D. M. Rudkevich,J. Rebek,Jr.,Chem. Eur. J.<br />

2000, 6,1007 – 1016; b) T. Haino,D. M. Rudkevich,A. Shivanyuk,K.<br />

Rissanen,J. Rebek,Jr.,Chem. Eur. J. 2000, 6,3797 –<br />

3805.<br />

[118] a) D. M. Rudkevich,G. Hilmersson,J. Rebek,Jr.,J. Am. Chem.<br />

Soc. 1997, 119,9911 – 9912; b) D. M. Rudkevich,G. Hilmersson,J.<br />

Rebek,Jr.,J. Am. Chem. Soc. 1998, 120,12 216 – 12225;<br />

c) S. H. Ma,D. M. Rudkevich,J. Rebek,Jr.,Angew. Chem.<br />

1999, 111,2761 – 2764; Angew. Chem. Int. Ed. 1999, 38,2600 –<br />

2602.<br />

[119] a) P. Amrhein,P. L. Wash,A. Shivanyuk,J. Rebek,Jr.,Org.<br />

Lett. 2002, 4,319 – 321; b) P. Amrhein,A. Shivanyuk,D. W.<br />

Johnson,J. Rebek,Jr.,J. Am. Chem. Soc. 2002, 124,10349 –<br />

10358.<br />

[120] a) C. D. Gutsche, Calixarenes: Monographs in Supramolecular<br />

Chemistry, Vol. 1,RSC,Cambridge,1989; b)Calixarenes: A<br />

Versatile Class of Macrocyclic Compounds: Topics in Inclusion<br />

Science, Vol. 3 (Eds.: J. Vicens,V. Böhmer),Kluwer Academic,<br />

Dordrecht, 1991.<br />

[121] For reviews,see: a) A. Ikeda,S. Shinkai,Chem. Rev. 1997, 97,<br />

1713 – 1734; b) S. Shinkai,A. Ikeda,Pure Appl. Chem. 1999, 71,<br />

275 – 280.<br />

[122] P. R. Ashton,R. Ballardini,V. Balzani,S. E. Boyd,A. Credi,<br />

M. T. G<strong>and</strong>olfi,M. Gómez-López,S. Iqbal,D. Philp,J. A.<br />

Preece,L. Prodi,H. G. Ricketts,J. F. Stoddart,M. S. Tolley,M.<br />

Venturi,A. J. P. White,D. J. Williams,Chem. Eur. J. 1997, 3,<br />

152 – 170.<br />

[123] Similar conformational changes have been observed in a<br />

number of related analogues,see: a) P. R. Ashton,M.<br />

Gómez-López,S. Iqbal,J. A. Preece,J. F. Stoddart,Tetrahedron<br />

Lett. 1997, 38,3635 – 3638; b) M. B. Nielsen,S. B. Nielsen,J.<br />

Becher, Chem. Commun. 1998,475 – 476; c) M. B. Nielsen,J. G.<br />

Hansen,J. Becher,Eur. J. Org. Chem. 1999,2807 – 2815; d) V.<br />

Balzani,P. Ceroni,A. Credi,M. Gómez-López,C. Hamers,J. F.<br />

Stoddart,R. Wolf,New J. Chem. 2001, 25,25 – 31; e) Y. Liu,<br />

A. H. Flood,J. F. Stoddart,J. Am. Chem. Soc. 2004, 126,9150 –<br />

9151.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

171


Reviews<br />

[124] See,for example: a) F. C. de Schryver,P. Collart,J. V<strong>and</strong>endriessche,R.<br />

Goedeweeck,A. Swinnen,M. van der Auweraer,<br />

Acc. Chem. Res. 1987, 20,159 – 166; b) H. Itagaki,W. Masuda,<br />

Y. Hirayanagi, Chem. Phys. Lett. 1999, 309,402 – 406; c) H.<br />

Itagaki,W. Masuda,Y. Hirayanagi,S. Sugimoto,J. Phys. Chem.<br />

B 2002, 106,3316 – 3322.<br />

[125] See,for example: a) X. Y. Lauteslager,M. J. Bartels,J. J. Piet,<br />

J. M. Warman,J. W. Verhoeven,A. M. Brouwer,Eur. J. Org.<br />

Chem. 1998,2467 – 2481; b) X. Y. Lauteslager,I. H. M. van<br />

Stokkum,H. J. van Ramesdonk,A. M. Brouwer,J. W. Verhoeven,<br />

J. Phys. Chem. A 1999, 103,653 – 659; c) T. D. M. Bell,<br />

K. A. Jolliffe,K. P. Ghiggino,A. M. Oliver,M. J. Shephard,S. J.<br />

Langford,M. N. Paddon-Row,J. Am. Chem. Soc. 2000, 122,<br />

10661 – 10666; d) X. Y. Lauteslager,I. H. M. van Stokkum,<br />

H. J. van Ramesdonk,D. Bebelaar,J. Fraanje,K. Goubitz,H.<br />

Schenk,A. M. Brouwer,J. W. Verhoeven,Eur. J. Org. Chem.<br />

2001,3105 – 3118.<br />

[126] For reviews <strong>and</strong> highlights with general coverage,see: a) A.<br />

Williams, Chem. Eur. J. 1997, 3,15 – 19; b) S. H. Gellman,Acc.<br />

Chem. Res. 1998, 31,173 – 180; c) A. E. Rowan,R. J. M. Nolte,<br />

Angew. Chem. 1998, 110,65 – 71; Angew. Chem. Int. Ed. 1998,<br />

37,63 – 68; d) M. S. Cubberley,B. L. Iverson,Curr. Opin. Chem.<br />

Biol. 2001, 5,650 – 653; e) D. J. Hill,M. J. Mio,R. B. Prince,<br />

T. S. Hughes,J. S. Moore,Chem. Rev. 2001, 101,3893 – 4011;<br />

f) C. Schmuck, Angew. Chem. 2003, 115,2552 – 2556; Angew.<br />

Chem. Int. Ed. 2003, 42,2448 – 2452; g) M. Albrecht,Angew.<br />

Chem. 2005, 117,6606 – 6609; Angew. Chem. Int. Ed. 2005, 44,<br />

6448 – 6451; on aromatic oligoamides: h) B. Gong, Chem. Eur.<br />

J. 2001, 7,4336 – 4342; i) A. R. Sanford,K. Yamato,X. W. Yang,<br />

L. H. Yuan,Y. H. Han,B. Gong,Eur. J. Biochem. 2004, 271,<br />

1416 – 1425; j) I. Huc, Eur. J. Org. Chem. 2004,17 – 29; on<br />

metal-templated helicates: k) E. C. Constable, Tetrahedron<br />

1992, 48,10013 – 10 059; l) C. Piguet,G. Bernardinelli,G.<br />

Hopfgartner, Chem. Rev. 1997, 97,2005 – 2062; m) M. Albrecht,<br />

Chem. Rev. 2001, 101,3457 – 3497; on biomimetic foldamers,for<br />

example, b-peptides <strong>and</strong> peptide nucleic acid: n) D. Seebach,<br />

J. L. Matthews, Chem. Commun. 1997,2015 – 2022; o) P. E.<br />

Nielsen,G. Haaima,Chem. Soc. Rev. 1997, 26,73 – 78; p) K. D.<br />

Stigers,M. J. Soth,J. S. Nowick,Curr. Opin. Chem. Biol. 1999,<br />

3,714 – 723; q) J. S. Nowick,Acc. Chem. Res. 1999, 32,287 –<br />

296; r) P. E. Nielsen, Acc. Chem. Res. 1999, 32,624 – 630; s) K.<br />

Kirshenbaum,R. N. Zuckermann,K. A. Dill,Curr. Opin.<br />

Struct. Biol. 1999, 9,530 – 535; t) P. Herdewijn,Biochim.<br />

Biophys. Acta 1999, 1489,167 – 179; u) R. P. Cheng,S. H.<br />

Gellman,W. F. DeGrado,Chem. Rev. 2001, 101,3219 – 3232;<br />

v) J. A. Patch,A. E. Barron,Curr. Opin. Chem. Biol. 2002, 6,<br />

872 – 877; w) T. A. Martinek,F. Fülöp,Eur. J. Biochem. 2003,<br />

270,3657 – 3666; x) R. P. Cheng,Curr. Opin. Struct. Biol. 2004,<br />

14,512 – 520; y) G. Licini,L. J. Prins,P. Scrimin,Eur. J. Org.<br />

Chem. 2005,969 – 977.<br />

[127] a) J. C. Nelson,J. G. Saven,J. S. Moore,P. G. Wolynes,Science<br />

1997, 277,1793 – 1796; b) R. B. Prince,J. G. Saven,P. G.<br />

Wolynes,J. S. Moore,J. Am. Chem. Soc. 1999, 121,3114 –<br />

3121; c) S. Lahiri,J. L. Thompson,J. S. Moore,J. Am. Chem.<br />

Soc. 2000, 122,11 315 – 11319; d) D. J. Hill,J. S. Moore,Proc.<br />

Natl. Acad. Sci. USA 2002, 99,5053 – 5057; derivatization with<br />

suitable coordinating units allows metal ions to assist in the<br />

helix formation,see: e) R. B. Prince,T. Okada,J. S. Moore,<br />

Angew. Chem. 1999, 111,245 – 249; Angew. Chem. Int. Ed.<br />

1999, 38,233 – 236.<br />

[128] For solvent-induced switching of helical screw-sense in a chiral<br />

helical structure,see: A. L. Hofacker,J. R. Parquette,Angew.<br />

Chem. 2005, 117,1077 – 1081; Angew. Chem. Int. Ed. 2005, 44,<br />

1053 – 1057.<br />

[129] a) V. Berl,I. Huc,R. G. Khoury,M. J. Krische,J.-M. Lehn,<br />

Nature 2000, 407,720 – 723; b) V. Berl,I. Huc,R. G. Khoury,J.-<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

M. Lehn, Chem. Eur. J. 2001, 7,2798 – 2809; c) V. Berl,I. Huc,<br />

R. G. Khoury,J.-M. Lehn,Chem. Eur. J. 2001, 7,2810 – 2820.<br />

[130] C. Dolain,V. Maurizot,I. Huc,Angew. Chem. 2003, 115,2844 –<br />

2846; Angew. Chem. Int. Ed. 2003, 42,2738 – 2740.<br />

[131] E. Kolomiets,V. Berl,I. Odriozola,A. M. Stadler,N. Kyritsakas,J.-M.<br />

Lehn,Chem. Commun. 2003,2868 – 2869.<br />

[132] D. Kanamori,T. A. Okamura,H. Yamamoto,N. Ueyama,<br />

Angew. Chem. 2005, 117,991 – 994; Angew. Chem. Int. Ed.<br />

2005, 44,969 – 972.<br />

[133] a) G. S. Hanan,J.-M. Lehn,N. Kyritsakas,J. Fischer,J. Chem.<br />

Soc. Chem. Commun. 1995,765 – 766; b) D. M. Bassani,J.-M.<br />

Lehn,G. Baum,D. Fenske,Angew. Chem. 1997, 109,1931 –<br />

1933; Angew. Chem. Int. Ed. Engl. 1997, 36,1845 – 1847; c) M.<br />

Ohkita,J.-M. Lehn,G. Baum,D. Fenske,Chem. Eur. J. 1999, 5,<br />

3471 – 3481.<br />

[134] M. Barboiu,J.-M. Lehn,Proc. Natl. Acad. Sci. USA 2002, 99,<br />

5201 – 5206.<br />

[135] A. M. Stadler,N. Kyritsakas,J.-M. Lehn,Chem. Commun. 2004,<br />

2024 – 2025.<br />

[136] M. Barboiu,G. Vaughan,N. Kyritsakas,J.-M. Lehn,Chem. Eur.<br />

J. 2003, 9,763 – 769.<br />

[137] A. Petitjean,R. G. Khoury,N. Kyritsakas,J.-M. Lehn,J. Am.<br />

Chem. Soc. 2004, 126,6637 – 6647.<br />

[138] F. Zhang,S. Bai,G. P. A. Yap,V. Tarwade,J. M. Fox,J. Am.<br />

Chem. Soc. 2005, 127,10 590 – 10599.<br />

[139] a) R. B. Woodward,R. Hoffmann,Angew. Chem. 1969, 81,<br />

797 – 869; Angew. Chem. Int. Ed. Engl. 1969, 8,781 – 853; b) K.<br />

Fukui, Acc. Chem. Res. 1971, 4,57 – 64; c) J. B. Hendrickson,<br />

Angew. Chem. 1974, 86,71 – 100; Angew. Chem. Int. Ed. Engl.<br />

1974, 13,47 – 76; d) W. R. Dolbier,H. Koroniak,K. N. Houk,<br />

C. M. Sheu, Acc. Chem. Res. 1996, 29,471 – 477.<br />

[140] a) B. E. R. Schilling,R. Hoffmann,J. Am. Chem. Soc. 1978, 100,<br />

6274 – 6274; b) B. E. R. Schilling,R. Hoffmann,J. Am. Chem.<br />

Soc. 1979, 101,3456 – 3467; c) R. T. Edidin,J. R. Norton,K.<br />

Mislow, Organometallics 1982, 1,561 – 562.<br />

[141] H. Ushiyama,K. Takatsuka,Angew. Chem. 2005, 117,1263 –<br />

1266; Angew. Chem. Int. Ed. 2005, 44,1237 – 1240.<br />

[142] There are differences between the technical definition of the<br />

terms “configuration” <strong>and</strong> “conformation” <strong>and</strong> their common<br />

usage. While formally,configurational differences involve<br />

differences in bond angles <strong>and</strong> conformational differences<br />

suggest changes in torsion angles,here “configuration” is<br />

applied to cis–trans isomerism around double bonds as<br />

convention suggests. The same categorization can be extended<br />

to include isomerism around the C N bond of amides,although<br />

this is often considered as a conformational change around a<br />

formal single bond. For further discussion,see Ref. [107]. The<br />

arrangement in space of the components of interlocked<br />

molecules <strong>and</strong> supramolecular complexes requires definition<br />

of another stereochemical term: “co-conformation”. This refers<br />

to the relative positions of the noncovalently linked components<br />

with respect to each other,see: M. C. T. Fyfe,P. T. Glink,<br />

S. Menzer,J. F. Stoddart,A. J. P. White,D. J. Williams,Angew.<br />

Chem. 1997, 109,2158 – 2160; Angew. Chem. Int. Ed. Engl.<br />

1997, 36,2068 – 2070.<br />

[143] For a comprehensive review of isomerizations in both natural<br />

<strong>and</strong> synthetic systems,see: C. Dugave,L. Demange,Chem. Rev.<br />

2003, 103,2475 – 2532.<br />

[144] a) D. H. Waldeck, Chem. Rev. 1991, 91,415 – 436; b) H. Meier,<br />

Angew. Chem. 1992, 104,1425 – 1446; Angew. Chem. Int. Ed.<br />

Engl. 1992, 31,1399 – 1420; c) J. Saltiel,D. F. Sears,Jr.,D.-H.<br />

Ko,K.-M. Park in CRC H<strong>and</strong>book of Organic Photochemistry<br />

<strong>and</strong> Photobiology (Eds.: W. M. Horspool,P.-S. Song),CRC,<br />

Boca Raton, 1995,pp. 3 – 15.<br />

[145] A photochromic process is one in which a light-induced<br />

interconversion between two species results in a change in<br />

absorption spectrum.<br />

172 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

[146] H. Suginome in CRC H<strong>and</strong>book of Organic Photochemistry<br />

<strong>and</strong> Photobiology (Eds.: W. M. Horspool,P.-S. Song),CRC,<br />

Boca Raton, 1995,pp. 824 – 840.<br />

[147] a) M. Irie, Pure Appl. Chem. 1996, 68,1367 – 1371; b) M. Irie,<br />

Chem. Rev. 2000, 100,1685 – 1716.<br />

[148] a) H. G. Heller in CRC H<strong>and</strong>book of Organic Photochemistry<br />

<strong>and</strong> Photobiology (Eds.: W. M. Horspool,P.-S. Song),CRC,<br />

Boca Raton, 1995,pp. 173 – 183; b) Y. Yokoyama,Chem. Rev.<br />

2000, 100,1717 – 1739; c) Y. Yokoyama in <strong>Molecular</strong> Switches<br />

(Ed.: B. L. Feringa),Wiley-VCH,Weinheim,2001,pp. 107 –<br />

121.<br />

[149] G. Berkovic,V. Krongauz,V. Weiss,Chem. Rev. 2000, 100,<br />

1741 – 1753.<br />

[150] a) B. L. Feringa,N. P. M. Huck,A. M. Schoevaars,Adv. Mater.<br />

1996, 8,681 – 684; b) B. L. Feringa,R. A. van Delden,N.<br />

Koumura,E. M. Geertsema,Chem. Rev. 2000, 100,1789 –<br />

1816; c) B. L. Feringa,R. A. van Delden,M. K. J. Ter Wiel<br />

in <strong>Molecular</strong> Switches (Ed.: B. L. Feringa),Wiley-VCH,<br />

Weinheim, 2001,pp. 123 – 163.<br />

[151] For reviews,see: a) B. L. Feringa,W. F. Jager,B. de Lange,<br />

Tetrahedron 1993, 49,8267 – 8310; b) special issue on “Photochromism:<br />

Memories <strong>and</strong> Switches”: Chem. Rev. 2000, 100,<br />

1683 – 1890; c) <strong>Molecular</strong> Switches (Ed.: B. L. Feringa),Wiley-<br />

VCH,Weinheim,2001.<br />

[152] a) G. Fischer, Angew. Chem. 1994, 106,1479 – 1501; Angew.<br />

Chem. Int. Ed. Engl. 1994, 33,1415 – 1436; b) G. Fischer,Chem.<br />

Soc. Rev. 2000, 29,119 – 127; c) J. Balbach,F. X. Schmid in<br />

Mechanisms of Protein Folding: Frontiers in <strong>Molecular</strong> Biology,<br />

Vol. 32 (Ed.: R. H. Pain),2nd ed.,Oxford University Press,<br />

Oxford, 2000,pp. 212 – 249.<br />

[153] C. Tie,J. C. Gallucci,J. R. Parquette,J. Am. Chem. Soc. 2006,<br />

128,1162 – 1171.<br />

[154] For reviews,see: a) A. Momotake,T. Arai,J. Photochem.<br />

Photobiol. C 2004, 5,1 – 25; b) A. Momotake,T. Arai,Polymer<br />

2004, 45,5369 – 5390.<br />

[155] T. Muraoka,K. Kinbara,Y. Kobayashi,T. Aida,J. Am. Chem.<br />

Soc. 2003, 125,5612 – 5613.<br />

[156] Multiple azobenzene chromophores have previously been used<br />

to cause significant geometric changes in macrocyclic <strong>and</strong><br />

macrobicyclic systems,see: a) S. Shinkai,Y. Honda,Y. Kusano,<br />

O. Manabe, J. Chem. Soc. Chem. Commun. 1982,848 – 850;<br />

b) M. Bauer,F. Vögtle,Chem. Ber. 1992, 125,1675 – 1686.<br />

[157] Y. Norikane,N. Tamaoki,Org. Lett. 2004, 6,2595 – 2598.<br />

[158] A. M. Schoevaars,W. Kruizinga,R. W. J. Zijlstra,N. Veldman,<br />

A. L. Spek,B. L. Feringa,J. Org. Chem. 1997, 62,4943 – 4948.<br />

[159] a) N. Harada,N. Koumura,B. L. Feringa,J. Am. Chem. Soc.<br />

1997, 119,7256 – 7264; b) N. Koumura,N. Harada,Chem. Lett.<br />

1998,1151 – 1152.<br />

[160] N. Koumura,R. W. J. Zijlstra,R. A. van Delden,N. Harada,<br />

B. L. Feringa, Nature 1999, 401,152 – 155.<br />

[161] For accounts on the development of the Feringa <strong>and</strong> Harada<br />

systems,see: a) B. L. Feringa,Acc. Chem. Res. 2001, 34,504 –<br />

513; b) B. L. Feringa,N. Koumura,R. A. van Delden,M. K. J.<br />

ter Wiel, Appl. Phys. A 2002, 75,301 – 308; c) B. L. Feringa,<br />

R. A. van Delden,M. K. J. ter Wiel,Pure Appl. Chem. 2003, 75,<br />

563 – 575.<br />

[162] The more facile thermal helix inversion for cis isomers of this<br />

first generation series of overcrowded alkene molecular motors<br />

appears to be a general trend,see: a) N. Harada,A. Saito,N.<br />

Koumura,D. C. Roe,W. F. Jager,R. W. J. Zijlstra,B. de Lange,<br />

B. L. Feringa, J. Am. Chem. Soc. 1997, 119,7249 – 7255;<br />

b) R. W. J. Zijlstra,W. F. Jager,B. de Lange,P. T. van Duijnen,<br />

B. L. Feringa,H. Goto,A. Saito,N. Koumura,N. Harada,J.<br />

Org. Chem. 1999, 64,1667 – 1674.<br />

[163] R. W. J. Zijlstra,P. T. van Duijnen,B. L. Feringa,T. Steffen,K.<br />

Duppen,D. A. Wiersma,J. Phys. Chem. A 1997, 101,9828 –<br />

9836.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

[164] M. K. J. ter Wiel,R. A. van Delden,A. Meetsma,B. L. Feringa,<br />

J. Am. Chem. Soc. 2005, 127,14208 – 14222.<br />

[165] a) N. Koumura,E. M. Geertsema,A. Meetsma,B. L. Feringa,J.<br />

Am. Chem. Soc. 2000, 122,12005 – 12006; b) N. Koumura,<br />

E. M. Geertsema,M. B. van Gelder,A. Meetsma,B. L. Feringa,<br />

J. Am. Chem. Soc. 2002, 124,5037 – 5051; c) E. M.<br />

Geertsema,N. Koumura,M. K. J. ter Wiel,A. Meetsma,B. L.<br />

Feringa, Chem. Commun. 2002,2962 – 2963; d) R. A. van Delden,N.<br />

Koumura,A. Schoevaars,A. Meetsma,B. L. Feringa,<br />

Org. Biomol. Chem. 2003, 1,33 – 35; e) D. Pijper,R. A.<br />

van Delden,A. Meetsma,B. L. Feringa,J. Am. Chem. Soc.<br />

2005, 127,17 612 – 17613.<br />

[166] a) M. K. J. ter Wiel,R. A. van Delden,A. Meetsma,B. L.<br />

Feringa, J. Am. Chem. Soc. 2003, 125,15076 – 15086. Based<br />

on comparison of CD data with earlier analogues,the absolute<br />

stereochemistry of the single diastereomer of 50 produced in<br />

this report was originally assigned as (2R,2’R)-(P,P)-trans-50.A<br />

recent enantioselective synthesis <strong>and</strong> stereochemical study by<br />

X-ray crystallography,however,has shown that this assignment<br />

should be revised to (2S,2’S)-(M,M)-trans-50,see: b) T. Fujita,<br />

S. Kuwahara,N. Harada,Eur. J. Org. Chem. 2005,4533 – 4543;<br />

of course,this only affects the absolute sense of rotation of the<br />

two halves of the motor. Isolation <strong>and</strong> analysis of the unstablecis<br />

isomer has been achieved,as has dectection of the unstabletrans<br />

form,thus indicating that rotation occurs through the<br />

same mechanism to the first <strong>and</strong> second generation systems,<br />

see: c) S. Kuwahara,T. Fujita,N. Harada,Eur. J. Org. Chem.<br />

2005,4544 – 4556. In this third generation system,however,it<br />

appears that the unstable-trans form is more thermally labile<br />

than the unstable-cis form,in contrast to the first generation<br />

system (see Refs. [166a,166c]).<br />

[167] For 50,it has recently been shown that the photostationary<br />

state ratios can be improved by more careful selection of<br />

irradiation wavelength,see Ref. [166c].<br />

[168] M. K. J. ter Wiel,R. A. van Delden,A. Meetsma,B. L. Feringa,<br />

Org. Biomol. Chem. 2005, 3,4071 – 4076.<br />

[169] a) J.-M. Lehn, Angew. Chem. 1988, 100,91 – 116; Angew. Chem.<br />

Int. Ed. Engl. 1988, 27,89 – 112; b) D. J. Cram,Angew. Chem.<br />

1988, 100,1041 – 1052; Angew. Chem. Int. Ed. Engl. 1988, 27,<br />

1009 – 1020; c) C. J. Pedersen, Angew. Chem. 1988, 100,1053 –<br />

1059; Angew. Chem. Int. Ed. Engl. 1988, 27,1021 – 1027; d) J.-<br />

M. Lehn, Supramolecular Chemistry: Concepts <strong>and</strong> Perspectives,VCH,Weinheim,1995;<br />

e)Comprehensive Supramolecular<br />

Chemistry (Eds.: J.-M. Lehn,J. L. Atwood,J. E. D. Davies,<br />

D. D. MacNicol,F. Vögtle),Pergamon,Oxford,1996; f) special<br />

issue on “<strong>Molecular</strong> Recognition”: Chem. Rev. 1997, 97,1231 –<br />

1734.<br />

[170] Myosin requires an actin track to move along: a) S. M. Block,<br />

Cell 1996, 87,151 – 157; while kinesin progresses along the<br />

pathway provided by microtubule filaments: b) S. M. Block,<br />

Cell 1998, 93,5 – 8; even ion pumps operate through precisely<br />

defined structural channels: c) J. C. Skou, Angew. Chem. 1998,<br />

110,2452 – 2461; Angew. Chem. Int. Ed. 1998, 37,2321 – 2328.<br />

[171] S. Shinkai,T. Nakaji,T. Ogawa,K. Shigematsu,O. Manabe,J.<br />

Am. Chem. Soc. 1981, 103,111 – 115.<br />

[172] S. Shinkai,M. Ishihara,K. Ueda,O. Manabe,J. Chem. Soc.<br />

Perkin Trans. 2 1985,511 – 518.<br />

[173] a) L. Fabbrizzi,M. Licchelli,P. Pallavicini,Acc. Chem. Res.<br />

1999, 32,846 – 853; b) V. Amendola,L. Fabbrizzi,C. Mangano,<br />

P. Pallavicini, Acc. Chem. Res. 2001, 34,488 – 493; c) V.<br />

Amendola,L. Fabbrizzi,C. Mangano,P. Pallavicini,Struct.<br />

Bonding (Berlin) 2001, 99,79 – 115.<br />

[174] L. Fabbrizzi,F. Gatti,P. Pallavicini,E. Zambarbieri,Chem. Eur.<br />

J. 1999, 5,682 – 690.<br />

[175] A similar anion translocation process has also been achieved in<br />

a related trinuclear,multicomponent system,see: G. DeSantis,<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

173


Reviews<br />

L. Fabbrizzi,D. Iacopino,P. Pallavicini,A. Perotti,A. Poggi,<br />

Inorg. Chem. 1997, 36,827 – 832.<br />

[176] A. Ikeda,T. Tsudera,S. Shinkai,J. Org. Chem. 1997, 62,3568 –<br />

3574.<br />

[177] L. Zelikovich,J. Libman,A. Shanzer,Nature 1995, 374,790 –<br />

792.<br />

[178] Metal-ion translocation between two binding sites has been<br />

achieved in a number of other systems based either on<br />

switching the oxidation state of the metal or on the protonation<br />

state of the receptor,see: Refs. [173b,173c] <strong>and</strong> a) E. Kimura,<br />

Y. Kurogi,T. Tojo,M. Shionoya,M. Shiro,J. Am. Chem. Soc.<br />

1991, 113,4857 – 4864; b) C. Belle,J. L. Pierre,E. Saint-Aman,<br />

New J. Chem. 1998, 22,1399 – 1402; c) T. R. Ward,A. Lutz,S. P.<br />

Parel,J. Ensling,P. Gütlich,P. Buglyo,C. Orvig,Inorg. Chem.<br />

1999, 38,5007 – 5017; d) V. Amendola,L. Fabbrizzi,C. Mangano,P.<br />

Pallavicini,A. Perotti,A. Taglietti,J. Chem. Soc.<br />

Dalton Trans. 2000,185 – 189; e) D. Kalny,M. Elhabiri,T.<br />

Moav,A. Vaskevich,I. Rubinstein,A. Shanzer,A.-M.<br />

Albrecht-Gary, Chem. Commun. 2002,1426 – 1427; f) V. Amendola,L.<br />

Fabbrizzi,C. Mangano,H. Miller,P. Pallavicini,A.<br />

Perotti,A. Taglietti,Angew. Chem. 2002, 114,2665 – 2668;<br />

Angew. Chem. Int. Ed. 2002, 41,2553 – 2556; g) L. Fabbrizzi,F.<br />

Foti,S. Patroni,P. Pallavicini,A. Taglietti,Angew. Chem. 2004,<br />

116,5183 – 5187; Angew. Chem. Int. Ed. 2004, 43,5073 – 5077.<br />

[179] For representative examples of pseudorotaxane architectures<br />

based on polyether macrocycles bound to nitrogen-based<br />

cationic linear components,see: a) P. R. Ashton,P. J. Campbell,<br />

E. J. T. Chrystal,P. T. Glink,S. Menzer,D. Philp,N. Spencer,<br />

J. F. Stoddart,P. A. Tasker,D. J. Williams,Angew. Chem. 1995,<br />

107,1997 – 2001; Angew. Chem. Int. Ed. Engl. 1995, 34,1865 –<br />

1869; b) P. R. Ashton,E. J. T. Chrystal,P. T. Glink,S. Menzer,<br />

C. Schiavo,J. F. Stoddart,P. A. Tasker,D. J. Williams,Angew.<br />

Chem. 1995, 107,2001 – 2004; Angew. Chem. Int. Ed. Engl.<br />

1995, 34,1869 – 1871; c) P. R. Ashton,E. J. T. Chrystal,P. T.<br />

Glink,S. Menzer,C. Schiavo,N. Spencer,J. F. Stoddart,P. A.<br />

Tasker,A. J. P. White,D. J. Williams,Chem. Eur. J. 1996, 2,<br />

709 – 728; d) P. R. Ashton,M. C. T. Fyfe,P. T. Glink,S. Menzer,<br />

J. F. Stoddart,A. J. P. White,D. J. Williams,J. Am. Chem. Soc.<br />

1997, 119,12 514 – 12524; e) S. J. Loeb,J. A. Wisner,Angew.<br />

Chem. 1998, 110,3010 – 3013; Angew. Chem. Int. Ed. 1998, 37,<br />

2838 – 2840; f) P. R. Ashton,I. Baxter,M. C. T. Fyfe,F. M.<br />

Raymo,N. Spencer,J. F. Stoddart,A. J. P. White,D. J. Williams,<br />

J. Am. Chem. Soc. 1998, 120,2297 – 2307; g) P. R. Ashton,<br />

M. C. T. Fyfe,M.-V. Martínez-Díaz,S. Menzer,C. Schiavo,J. F.<br />

Stoddart,A. J. P. White,D. J. Williams,Chem. Eur. J. 1998, 4,<br />

1523 – 1534; h) T. Clifford,A. Abushamleh,D. H. Busch,Proc.<br />

Natl. Acad. Sci. USA 2002, 99,4830 – 4836; i) H. W. Gibson,N.<br />

Yamaguchi,J. W. Jones,J. Am. Chem. Soc. 2003, 125,3522 –<br />

3533; j) S. J. Loeb,J. Tiburcio,S. J. Vella,Org. Lett. 2005, 7,<br />

4923 – 4926.<br />

[180] For examples of pseudorotaxane architectures based on pdonor<br />

guest <strong>and</strong> p-acceptor cyclophane interactions,see:<br />

a) P. L. Anelli,P. R. Ashton,N. Spencer,A. M. Z. Slawin,J. F.<br />

Stoddart,D. J. Williams,Angew. Chem. 1991, 103,1052 – 1054;<br />

Angew. Chem. Int. Ed. Engl. 1991, 30,1036 – 1039; b) L.<br />

Grubert,D. Jacobi,K. Buck,W. Abraham,C. Mugge,E.<br />

Krause, Eur. J. Org. Chem. 2001,3921 – 3932; c) J. O. Jeppesen,<br />

J. Becher,J. F. Stoddart,Org. Lett. 2002, 4,557 – 560; d) W.<br />

Abraham,L. Grubert,S. Schmidt-Schäffer,K. Buck,Eur. J.<br />

Org. Chem. 2005,374 – 386.<br />

[181] For examples of other types of pseudorotaxane architectures,<br />

see: Refs. [50b,c,f,i] <strong>and</strong> a) A. Mirzoian, A. E. Kaifer, Chem.<br />

Eur. J. 1997, 3,1052 – 1058; b) H. Yonemura,S. Kusano,T.<br />

Matsuo,S. Yamada,Tetrahedron Lett. 1998, 39,6915 – 6918;<br />

c) K. Eliadou,K. Yannakopoulou,A. Rontoyianni,I. M.<br />

Mavridis, J. Org. Chem. 1999, 64,6217 – 6226; d) K. M. Huh,<br />

T. Ooya,S. Sasaki,N. Yui,Macromolecules 2001, 34,2402 –<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

2404; e) J. A. Wisner,P. D. Beer,M. G. B. Drew,Angew. Chem.<br />

2001, 113,3718 – 3721; Angew. Chem. Int. Ed. 2001, 40,3606 –<br />

3609; f) J. A. Wisner,P. D. Beer,N. G. Berry,B. Tomapatanaget,<br />

Proc. Natl. Acad. Sci. USA 2002, 99,4983 – 4986; g) A.<br />

Arduini,F. Calzavacca,A. Pochini,A. Secchi,Chem. Eur. J.<br />

2003, 9,793 – 799; h) F. H. Huang,F. R. Fronczek,H. W.<br />

Gibson, J. Am. Chem. Soc. 2003, 125,9272 – 9273; i) F. H.<br />

Huang,H. W. Gibson,W. S. Bryant,D. S. Nagvekar,F. R.<br />

Fronczek, J. Am. Chem. Soc. 2003, 125,9367 – 9371; j) Y. Inoue,<br />

T. Kanbara,T. Yamamoto,Tetrahedron Lett. 2004, 45,4603 –<br />

4606; k) M. R. Sambrook,P. D. Beer,J. A. Wisner,R. L. Paul,<br />

A. R. Cowley,F. Szemes,M. G. B. Drew,J. Am. Chem. Soc.<br />

2005, 127,2292 – 2302; l) B. A. Blight,K. A. Van Noortwyk,<br />

J. A. Wisner,M. C. Jennings,Angew. Chem. 2005, 117,1523 –<br />

1528; Angew. Chem. Int. Ed. 2005, 44,1499 – 1504; m) T. Kraus,<br />

M. Buděsˇínsky´,J. C. Cva›ka,J.-P. Sauvage,Angew. Chem. 2006,<br />

118,264 – 267; Angew. Chem. Int. Ed. 2006, 45,258 – 261;<br />

n) M. V. Rekharsky,H. Yamamura,M. Kawai,I. Osaka,R.<br />

Arakawa,A. Sato,Y. H. Ko,N. Selvapalam,K. Kim,Y. Inoue,<br />

Org. Lett. 2006, 8,815 – 818.<br />

[182] The IUPAC Compendium of Chemical Terminology (A. D.<br />

McNaught,A. Wilkinson,2nd ed.,Blackwell Science,1997)<br />

defines rotaxanes as “molecules in which a ring encloses<br />

another,rod-like molecule having end groups too large to pass<br />

through the ring opening,<strong>and</strong> thus holds the rod-like molecule<br />

in position without covalent bonding.”. This definition formally<br />

excludes threaded host–guest complexes (pseudorotaxanes).<br />

[183] For a brief account of a number of such systems,see: V. Balzani,<br />

A. Credi,M. Venturi,Proc. Natl. Acad. Sci. USA 2002, 99,<br />

4814 – 4817.<br />

[184] a) R. Ballardini,V. Balzani,M. T. G<strong>and</strong>olfi,L. Prodi,M.<br />

Venturi,D. Philp,H. G. Ricketts,J. F. Stoddart,Angew. Chem.<br />

1993, 105,1362 – 1364; Angew. Chem. Int. Ed. Engl. 1993, 32,<br />

1301 – 1303; b) M. Montalti,R. Ballardini,L. Prodi,V. Balzani,<br />

Chem. Commun. 1996,2011 – 2012; c) R. Ballardini,V. Balzani,<br />

A. Credi,M. T. G<strong>and</strong>olfi,S. J. Langford,S. Menzer,L. Prodi,<br />

J. F. Stoddart,M. Venturi,D. J. Williams,Angew. Chem. 1996,<br />

108,1056 – 1059; Angew. Chem. Int. Ed. Engl. 1996, 35,978 –<br />

981; d) M. Asakawa,P. R. Ashton,V. Balzani,A. Credi,G.<br />

Mattersteig,O. A. Matthews,M. Montalti,N. Spencer,J. F.<br />

Stoddart,M. Venturi,Chem. Eur. J. 1997, 3,1992 – 1996; e) A.<br />

Credi,V. Balzani,S. J. Langford,J. F. Stoddart,J. Am. Chem.<br />

Soc. 1997, 119,2679 – 2681; f) P. R. Ashton,R. Ballardini,V.<br />

Balzani,M. Gómez-López,S. E. Lawrence,M.-V. Martínez-<br />

Díaz,M. Montalti,A. Piersanti,L. Prodi,J. F. Stoddart,D. J.<br />

Williams, J. Am. Chem. Soc. 1997, 119,10641 – 10651; g) W.<br />

Devonport,M. A. Blower,M. R. Bryce,L. M. Goldenberg,J.<br />

Org. Chem. 1997, 62,885 – 887; h) M. Montalti,L. Prodi,Chem.<br />

Commun. 1998,1461 – 1462; i) O. A. Matthews,F. M. Raymo,<br />

J. F. Stoddart,A. J. P. White,D. J. Williams,New J. Chem. 1998,<br />

22,1131 – 1134; j) M. Asakawa,P. R. Ashton,V. Balzani,S. E.<br />

Boyd,A. Credi,G. Mattersteig,S. Menzer,M. Montalti,F. M.<br />

Raymo,C. Ruffilli,J. F. Stoddart,M. Venturi,D. J. Williams,<br />

Eur. J. Org. Chem. 1999,985 – 994; k) M. Asakawa,P. R.<br />

Ashton,V. Balzani,C. L. Brown,A. Credi,O. A. Matthews,<br />

S. P. Newton,F. M. Raymo,A. N. Shipway,N. Spencer,A.<br />

Quick,J. F. Stoddart,A. J. P. White,D. J. Williams,Chem. Eur.<br />

J. 1999, 5,860 – 875; l) V. Balzani,A. Credi,G. Mattersteig,<br />

O. A. Matthews,F. M. Raymo,J. F. Stoddart,M. Venturi,<br />

A. J. P. White,D. J. Williams,J. Org. Chem. 2000, 65,1924 –<br />

1936; m) V. Balzani,J. Becher,A. Credi,M. B. Nielsen,F. M.<br />

Raymo,J. F. Stoddart,A. M. Talarico,M. Venturi,J. Org.<br />

Chem. 2000, 65,1947 – 1956; n) V. Balzani,A. Credi,F.<br />

Marchioni,J. F. Stoddart,Chem. Commun. 2001,1860 – 1861;<br />

o) T. Fujimoto,A. Nakamura,Y. Inoue,Y. Sakata,T. Kaneda,<br />

Tetrahedron Lett. 2001, 42,7987 – 7989; p) M. Horie,Y. Suzaki,<br />

K. Osakada, J. Am. Chem. Soc. 2004, 126,3684 – 3685; q) A.<br />

174 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Credi,S. Dumas,S. Silvi,M. Venturi,A. Arduini,A. Pochini,A.<br />

Secchi, J. Org. Chem. 2004, 69,5881 – 5887; r) K. J. Chang,Y. J.<br />

An,H. Uh,K. S. Jeong,J. Org. Chem. 2004, 69,6556 – 6563;<br />

s) G. Kaiser,T. Jarrosson,S. Otto,Y. F. Ng,A. D. Bond,J. K. M.<br />

S<strong>and</strong>ers, Angew. Chem. 2004, 116,1993 – 1996; Angew. Chem.<br />

Int. Ed. 2004, 43,1959 – 1962; t) M. Venturi,S. Dumas,V.<br />

Balzani,J. G. Cao,J. F. Stoddart,New J. Chem. 2004, 28,1032 –<br />

1037; u) S. Nygaard,B. W. Laursen,A. H. Flood,C. N. Hansen,<br />

J. O. Jeppesen,J. F. Stoddart,Chem. Commun. 2006,144 – 146.<br />

[185] a) M. Asakawa,S. Iqbal,J. F. Stoddart,N. D. Tinker,Angew.<br />

Chem. 1996, 108,1053 – 1056; Angew. Chem. Int. Ed. Engl.<br />

1996, 35,976 – 978; b) A. Credi,M. Montalti,V. Balzani,S. J.<br />

Langford,F. M. Raymo,J. F. Stoddart,New J. Chem. 1998, 22,<br />

1061 – 1065; c) P. R. Ashton,V. Balzani,J. Becher,A. Credi,<br />

M. C. T. Fyfe,G. Mattersteig,S. Menzer,M. B. Nielsen,F. M.<br />

Raymo,J. F. Stoddart,M. Venturi,D. J. Williams,J. Am. Chem.<br />

Soc. 1999, 121,3951 – 3957.<br />

[186] a) W. L. Mock,J. Pierpont,J. Chem. Soc. Chem. Commun.<br />

1990,1509 – 1511; b) S. I. Jun,J. W. Lee,S. Sakamoto,K.<br />

Yamaguchi,K. Kim,Tetrahedron Lett. 2000, 41,471 – 475;<br />

c) J. O. Jeppesen,S. A. Vignon,J. F. Stoddart,Chem. Eur. J.<br />

2003, 9,4611 – 4625.<br />

[187] a) P. R. Ashton,V. Balzani,O. Kocian,L. Prodi,N. Spencer,<br />

J. F. Stoddart, J. Am. Chem. Soc. 1998, 120,11 190 – 11191; b) E.<br />

Ishow,A. Credi,V. Balzani,F. Spadola,L. M<strong>and</strong>olini,Chem.<br />

Eur. J. 1999, 5,984 – 989; c) R. Ballardini,V. Balzani,M.<br />

Clemente-Leon,A. Credi,M. T. G<strong>and</strong>olfi,E. Ishow,J. Perkins,<br />

J. F. Stoddart,H.-R. Tseng,S. Wenger,J. Am. Chem. Soc. 2002,<br />

124,12786 – 12795.<br />

[188] For examples of mechanical devices based on kinetically stable<br />

pseudorotaxane systems,see: a) J.-P. Collin,P. Gaviæa,J.-P.<br />

Sauvage, Chem. Commun. 1996,2005 – 2006; b) J. W. Lee,K. P.<br />

Kim,K. Kim,Chem. Commun. 2001,1042 – 1043; c) W. S. Jeon,<br />

A. Y. Ziganshina,J. W. Lee,Y. H. Ko,J. K. Kang,C. Lee,K.<br />

Kim, Angew. Chem. 2003, 115,4231 – 4234; Angew. Chem. Int.<br />

Ed. 2003, 42,4097 – 4100; d) W. S. Jeon,E. Kim,Y. H. Ko,I. H.<br />

Hwang,J. W. Lee,S. Y. Kim,H. J. Kim,K. Kim,Angew. Chem.<br />

2005, 117,89 – 93; Angew. Chem. Int. Ed. 2005, 44,87 – 91.<br />

[189] a) G. Schill, Catenanes, Rotaxanes <strong>and</strong> Knots,Academic Press,<br />

New York, 1971; b) D. M. Walba, Tetrahedron 1985, 41,3161 –<br />

3212; c) D. B. Amabilino,J. F. Stoddart,Chem. Rev. 1995, 95,<br />

2725 – 2828; d) S. A. Nepogodiev,J. F. Stoddart,Chem. Rev.<br />

1998, 98,1959 – 1976; e) D. A. Leigh,A. Murphy,Chem. Ind.<br />

1999,178 – 183; f) G. A. Breault,C. A. Hunter,P. C. Mayers,<br />

Tetrahedron 1999, 55,5265 – 5293; g) <strong>Molecular</strong> Catenanes<br />

Rotaxanes <strong>and</strong> Knots (Eds.: J.-P. Sauvage,C. O. Dietrich-<br />

Buchecker),Wiley-VCH,Weinheim,1999.<br />

[190] There are some examples of catenanes <strong>and</strong> rotaxanes constructed<br />

under thermodynamic control by using dynamic<br />

processes such as olefin metathesis or formation of an imine<br />

bond (see for example: R. L. E. Furlan,S. Otto,J. K. M.<br />

S<strong>and</strong>ers, Proc. Natl. Acad. Sci. USA 2002, 99,4801 – 4804,<strong>and</strong><br />

references therein). Such systems cannot act as molecular<br />

machines if they exchange components with the bulk quicker<br />

than the timescale of their stimuli-induced motion.<br />

[191] For reviews on the development of interlocked molecules as<br />

molecular machines,see: Refs. [1c,h,m] <strong>and</strong> a) A. C. Benniston,<br />

Chem. Soc. Rev. 1996, 25,427 – 435; b) M. Gómez-López,J. A.<br />

Preece,J. F. Stoddart,Nanotechnology 1996, 7,183 – 192; c) V.<br />

Balzani,M. Gómez-López,J. F. Stoddart,Acc. Chem. Res. 1998,<br />

31,405 – 414; d) J.-P. Sauvage,Acc. Chem. Res. 1998, 31,611 –<br />

619; e) J.-C. Chambron,J.-P. Sauvage,Chem. Eur. J. 1998, 4,<br />

1362 – 1366; f) J.-P. Collin,P. Gaviæa,V. Heitz,J.-P. Sauvage,<br />

Eur. J. Inorg. Chem. 1998,1 – 14; g) M.-J. Blanco,M. C.<br />

JimØnez,J.-C. Chambron,V. Heitz,M. Linke,J.-P. Sauvage,<br />

Chem. Soc. Rev. 1999, 28,293 – 305; h) A. R. Pease,J. O.<br />

Jeppesen,J. F. Stoddart,Y. Luo,C. P. Collier,J. R. Heath,Acc.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Chem. Res. 2001, 34,433 – 444; i) A. Harada,Acc. Chem. Res.<br />

2001, 34,456 – 464; j) C. A. Schalley,K. Beizai,F. Vögtle,Acc.<br />

Chem. Res. 2001, 34,465 – 476; k) J.-P. Collin,C. Dietrich-<br />

Buchecker,P. Gaviæa,M. C. Jimenez-Molero,J.-P. Sauvage,<br />

Acc. Chem. Res. 2001, 34,477 – 487; l) L. Raehm,J.-P. Sauvage,<br />

Struct. Bonding (Berlin) 2001, 99,55 – 78; m) R. Ballardini,V.<br />

Balzani,A. Credi,M. T. G<strong>and</strong>olfi,M. Venturi,Struct. Bonding<br />

(Berlin) 2001, 99,163 – 188; n) A. R. Pease,J. F. Stoddart,<br />

Struct. Bonding (Berlin) 2001, 99,189 – 236; o) V. Balzani,A.<br />

Credi, Chem. Rec. 2001, 1,422 – 435; p) F. M. Raymo,J. F.<br />

Stoddart in <strong>Molecular</strong> Switches (Ed.: B. L. Feringa),Wiley-<br />

VCH,Weinheim,2001,pp. 219 – 248; q) J.-P. Collin,J.-M. Kern,<br />

L. Raehm,J.-P. Sauvage in <strong>Molecular</strong> Switches (Ed.: B. L.<br />

Feringa),Wiley-VCH,Weinheim,2001,pp. 249 – 280; r) B. X.<br />

Colasson,C. Dietrich-Buchecker,M. C. Jimenez-Molero,J.-P.<br />

Sauvage, J. Phys. Org. Chem. 2002, 15,476 – 483; s) V. Balzani,<br />

A. Credi,M. Venturi,Pure Appl. Chem. 2003, 75,541 – 547;<br />

t) A. M. Brouwer,S. M. Fazio,C. Frochot,F. G. Gatti,D. A.<br />

Leigh,J. K. Y. Wong,G. W. H. Wurpel,Pure Appl. Chem. 2003,<br />

75,1055 – 1060; u) C. Dietrich-Buchecker,M. C. Jimenez-<br />

Molero,V. Sartor,J.-P. Sauvage,Pure Appl. Chem. 2003, 75,<br />

1383 – 1393; v) M. C. Jimenez-Molero,C. Dietrich-Buchecker,<br />

J.-P. Sauvage, Chem. Commun. 2003,1613 – 1616; w) M. J.<br />

Gunter, Eur. J. Org. Chem. 2004,1655 – 1673; x) A. H. Flood,<br />

R. J. A. Ramirez,W.-Q. Deng,R. P. Muller,W. A. Goddard,<br />

J. F. Stoddart, Aust. J. Chem. 2004, 57,301 – 322; y) P. M.<br />

Mendes,A. H. Flood,J. F. Stoddart,Appl. Phys. A 2005, 80,<br />

1197 – 1209; z) J.-P. Sauvage, Chem. Commun. 2005,1507 –<br />

1510; aa) J.-P. Collin,J.-P. Sauvage,Chem. Lett. 2005, 34,<br />

742 – 747; ab) V. Balzani,A. Credi,B. Ferrer,S. Silvi,M.<br />

Venturi, Top. Curr. Chem. 2005, 262,1 – 27; ac) J.-P. Collin,V.<br />

Heitz,J.-P. Sauvage,Top. Curr. Chem. 2005, 262,29 – 62;<br />

ad) N. N. P. Moonen,A. H. Flood,J. M. Fernµndez,J. F. Stoddart,<br />

Top. Curr. Chem. 2005, 262,99 – 132; ae) E. R. Kay,D. A.<br />

Leigh, Top. Curr. Chem. 2005, 262,133 – 177; af) A. B.<br />

Braunschweig,B. H. Northrop,J. F. Stoddart,J. Mater. Chem.<br />

2006, 16,32 – 44.<br />

[192] Numbers in square brackets preceding the names of interlocked<br />

compounds indicate the number of mechanically<br />

interlocked species,for example,a [2]rotaxane consists of a<br />

single macrocycle locked onto one thread; two macrocycles on<br />

one thread would constitute a [3]rotaxane.<br />

[193] According to Prelog,the concept of a molecular catenane was<br />

discussed by Willstätters prior to 1912 (see Refs. [189a,b]). The<br />

first synthesis of a catenane (using statistical threading with no<br />

assisting template effects) was achieved by Wasserman in 1960:<br />

a) E. Wasserman, J. Am. Chem. Soc. 1960, 82,4433 – 4434;<br />

proving the viability of such interlocked <strong>and</strong> otherwise<br />

topologically interesting molecules: b) H. L. Frisch,E. Wasserman,<br />

J. Am. Chem. Soc. 1961, 83,3789 – 3795; the first directed<br />

synthesis of a catenane was reported by Schill <strong>and</strong> Lüttringhaus<br />

in 1964: c) G. Schill,A. Lüttringhaus,Angew. Chem. 1964, 76,<br />

567 – 568; Angew. Chem. Int. Ed. Engl. 1964, 3,546 – 547; the<br />

first rotaxane was synthesized (using statistical methods) by<br />

Harrison <strong>and</strong> Harrison in 1967: d) I. T. Harrison,S. Harrison,J.<br />

Am. Chem. Soc. 1967, 89,5723 – 5724; e) I. T. Harrison,J.<br />

Chem. Soc. Perkin Trans. 1 1974,301 – 304.<br />

[194] Another synthetic strategy,“slippage”,theoretically provides a<br />

route to rotaxanes under thermodynamic control. However,the<br />

chemical assemblies prepared thus far by this route clearly do<br />

not satisfy the structural requirements of a rotaxane since the<br />

components can be separated from each other without breaking<br />

covalent bonds. Rather they are pseudorotaxanes—<br />

threaded host–guest complexes—which are (reasonably) kinetically<br />

stable at room temperature. The misclassification (see<br />

Ref. [182]) is not normally an issue because the physical<br />

behavior of kinetically stable pseudorotaxanes generally mir-<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

175


Reviews<br />

rors that of rotaxanes. However,it increasingly appears to be<br />

confusing nonspecialists into thinking that rotaxanes are<br />

supramolecular species or that pseudorotaxanes that are not<br />

kinetically stable can act as molecular machines. Slippage could<br />

be used as a strategy to form rotaxanes if the stabilization<br />

gained from the ring binding on the thread was sufficient to<br />

mean that covalent bond breaking is less energetically dem<strong>and</strong>ing<br />

than dethreading of the rings over the stoppers. However,to<br />

date no structures of this type have been reported.<br />

[195] For general reviews on template synthesis,with particular<br />

regard to the development of methods for the assembly of<br />

mechanically interlocked species,see: Ref. [189] <strong>and</strong> a) C. O.<br />

Dietrich-Buchecker,J.-P. Sauvage,Chem. Rev. 1987, 87,795 –<br />

810; b) S. Anderson,H. L. Anderson,J. K. M. S<strong>and</strong>ers,Acc.<br />

Chem. Res. 1993, 26,469 – 475; c) R. Hoss,F. Vögtle,Angew.<br />

Chem. 1994, 106,389 – 398; Angew. Chem. Int. Ed. Engl. 1994,<br />

33,375 – 384; d) Templated Organic Synthesis (Eds.: F. Diederich,P.<br />

J. Stang),Wiley-VCH,Weinheim,1999; e)T.J.<br />

Hubin,D. H. Busch,Coord. Chem. Rev. 2000, 200,5–52; for<br />

particular accounts on the synthesis of interlocked architectures<br />

with metals as templates <strong>and</strong>/or as part the framework,see:<br />

f) J.-P. Sauvage, Acc. Chem. Res. 1990, 23,319 – 327; g) D. H.<br />

Busch,N. A. Stephenson,Coord. Chem. Rev. 1990, 100,119 –<br />

154; h) M. Fujita,K. Ogura,Coord. Chem. Rev. 1996, 148,249 –<br />

264; i) M. Fujita, Acc. Chem. Res. 1999, 32,53 – 61; j) C.<br />

Dietrich-Buchecker,G. Rapenne,J.-P. Sauvage,Coord. Chem.<br />

Rev. 1999, 185–186,167 – 176; k) J.-C. Chambron,J.-P. Collin,V.<br />

Heitz,D. Jouvenot,J.-M. Kern,P. Mobian,D. Pomeranc,J.-P.<br />

Sauvage, Eur. J. Org. Chem. 2004,1627 – 1638; l) S. J. Cantrill,<br />

K. S. Chichak,A. J. Peters,J. F. Stoddart,Acc. Chem. Res. 2005,<br />

38,1 – 9; see also: m) A.-M. Fuller,D. A. Leigh,P. J. Lusby,<br />

I. D. H. Oswald,S. Parsons,D. B. Walker,Angew. Chem. 2004,<br />

116,4004 – 4008; Angew. Chem. Int. Ed. 2004, 43,3914 – 3918;<br />

for accounts focusing on the synthesis of interlocked molecules<br />

templated by a combination of hydrogen bonds <strong>and</strong> p–<br />

p interactions,see: n) D. Philp,J. F. Stoddart,Synlett 1991,<br />

445 – 458; o) J. F. Stoddart, Chem. Br. 1991, 27,714 – 718;<br />

p) F. M. Raymo,J. F. Stoddart,Pure Appl. Chem. 1996, 68,<br />

313 – 322; q) R. E. Gillard,F. M. Raymo,J. F. Stoddart,Chem.<br />

Eur. J. 1997, 3,1933 – 1940; r) C. G. Claessens,J. F. Stoddart,J.<br />

Phys. Org. Chem. 1997, 10,254 – 272; s) F. M. Raymo,J. F.<br />

Stoddart, Chemtracts: Org. Chem. 1998, 11,491 – 511; t) P. T.<br />

Glink,J. F. Stoddart,Pure Appl. Chem. 1998, 70,419 – 424 (the<br />

latter also contains a brief account of templated synthesis based<br />

on ammonium ion/crown ether hydrogen bonds); for accounts<br />

of amide-based hydrogen bond directed synthesis of interlocked<br />

architectures,see: u) R. Jäger,F. Vögtle,Angew. Chem.<br />

1997, 109,966 – 980; Angew. Chem. Int. Ed. Engl. 1997, 36,930 –<br />

944; v) A. G. Johnston,D. A. Leigh,R. J. Pritchard,M. D.<br />

Deegan, Angew. Chem. 1995, 107,1324 – 1327; Angew. Chem.<br />

Int. Ed. Engl. 1995, 34,1209 – 1212; w) A. G. Johnston,D. A.<br />

Leigh,L. Nezhat,J. P. Smart,M. D. Deegan,Angew. Chem.<br />

1995, 107,1327 – 1331; Angew. Chem. Int. Ed. Engl. 1995, 34,<br />

1212 – 1216; x) A. G. Johnston,D. A. Leigh,A. Murphy,J. P.<br />

Smart,M. D. Deegan,J. Am. Chem. Soc. 1996, 118,10662 –<br />

10663; y) F. Biscarini,M. Cavallini,D. A. Leigh,S. León,S. J.<br />

Teat,J. K. Y. Wong,F. Zerbetto,J. Am. Chem. Soc. 2002, 124,<br />

225 – 233; z) V. Aucagne,D. A. Leigh,J. S. Lock,A. R. Thomson,<br />

J. Am. Chem. Soc. 2006, 128,1784 – 1785; for phenolate<br />

templates,see: aa) C. Seel,F. Vögtle,Chem. Eur. J. 2000, 6,21 –<br />

24; for a discussion of the template-directed construction of<br />

interlocked architectures under thermodynamic control,see<br />

Ref. [190] <strong>and</strong> ab) T. J. Kidd,D. A. Leigh,A. J. Wilson,J. Am.<br />

Chem. Soc. 1999, 121,1599 – 1600; ac) D. A. Leigh,P. J. Lusby,<br />

S. J. Teat,A. J. Wilson,J. K. Y. Wong,Angew. Chem. 2001, 113,<br />

1586 – 1591; Angew. Chem. Int. Ed. 2001, 40,1538 – 1543;<br />

ad) J. S. Hannam,T. J. Kidd,D. A. Leigh,A. J. Wilson,Org.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Lett. 2003, 5,1907 – 1910; ae) L. Hogg,D. A. Leigh,P. J. Lusby,<br />

A. Morelli,S. Parsons,J. K. Y. Wong,Angew. Chem. 2004, 116,<br />

1238 – 1241; Angew. Chem. Int. Ed. 2004, 43,1218 – 1221; for a<br />

route to rotaxanes that only requires a catalytic amount of<br />

template,see: af) V. Aucagne,K. D. Hänni,D. A. Leigh,P. J.<br />

Lusby,D. B. Walker,J. Am. Chem. Soc. 2006, 128,2186 – 2187.<br />

[196] With the notable exception of the earliest statistically constructed<br />

[2]rotaxanes (see Refs. [193d,e]) <strong>and</strong> those assembled<br />

by covalent bond-directed synthesis (a) G. Schill,H. Zollenkopf,<br />

Justus Liebigs Ann. Chem. 1969, 721,53 – 74; b) K.<br />

Hiratani,J. Suga,Y. Nagawa,H. Houjou,H. Tokuhisa,M.<br />

Numata,K. Watanabe,Tetrahedron Lett. 2002, 43,5747 – 5750;<br />

c) N. Kameta,A. Hiratani,Y. Nagawa,Chem. Commun. 2004,<br />

466 – 467; d) Y. Nagawa,J. Suga,K. Hiratani,E. Koyama,M.<br />

Kanesato, Chem. Commun. 2005,749 – 751),there are few<br />

examples of rotaxanes without significant noncovalent recognition<br />

elements between macrocycle <strong>and</strong> thread; however,see:<br />

Ref. [195aa,af] <strong>and</strong> e) G. M. Hübner, J. Glaser, C. Seel, F.<br />

Vögtle, Angew. Chem. 1999, 111,395 – 398; Angew. Chem. Int.<br />

Ed. 1999, 38,383 – 386; f) C. Reuter,W. Wien<strong>and</strong>,G. M.<br />

Hübner,C. Seel,F. Vögtle,Chem. Eur. J. 1999, 5,2692 – 2697;<br />

g) C. A. Schalley,G. Silva,C. F. Nising,P. Linnartz,Helv. Chim.<br />

Acta 2002, 85,1578 – 1596; h) J. M. Mahoney,R. Shukla,R. A.<br />

Marshall,A. M. Beatty,J. Zajicek,B. D. Smith,J. Org. Chem.<br />

2002, 67,1436 – 1440; i) X. Y. Li,J. Illigen,M. Nieger,S. Michel,<br />

C. A. Schalley, Chem. Eur. J. 2003, 9,1332 – 1347; j) J. S.<br />

Hannam,S. M. Lacy,D. A. Leigh,C. G. Saiz,A. M. Z. Slawin,<br />

S. G. Stitchell, Angew. Chem. 2004, 116,3322 – 3326; Angew.<br />

Chem. Int. Ed. 2004, 43,3260 – 3264.<br />

[197] P. L. Anelli,N. Spencer,J. F. Stoddart,J. Am. Chem. Soc. 1991,<br />

113,5131 – 5133.<br />

[198] a) P. R. Ashton,D. Philp,N. Spencer,J. F. Stoddart,J. Chem.<br />

Soc. Chem. Commun. 1992,1124 – 1128; b) P. R. Ashton,M. R.<br />

Johnston,J. F. Stoddart,M. S. Tolley,J. W. Wheeler,J. Chem.<br />

Soc. Chem. Commun. 1992,1128 – 1131.<br />

[199] A. S. Lane,D. A. Leigh,A. Murphy,J. Am. Chem. Soc. 1997,<br />

119,11092 – 11093.<br />

[200] D. A. Leigh,A. Murphy,J. P. Smart,A. M. Z. Slawin,Angew.<br />

Chem. 1997, 109,752 – 756; Angew. Chem. Int. Ed. Engl. 1997,<br />

36,728 – 732.<br />

[201] The rate of submolecular motion in the analogous benzylic<br />

amide catenanes exhibits similar solvent effects,see: a) D. A.<br />

Leigh,A. Murphy,J. P. Smart,M. S. Deleuze,F. Zerbetto,J.<br />

Am. Chem. Soc. 1998, 120,6458 – 6467; b) M. S. Deleuze,D. A.<br />

Leigh,F. Zerbetto,J. Am. Chem. Soc. 1999, 121,2364 – 2379;<br />

c) D. A. Leigh,A. Troisi,F. Zerbetto,Chem. Eur. J. 2001, 7,<br />

1450 – 1454.<br />

[202] P. Ghosh,G. Federwisch,M. Kogej,C. A. Schalley,D. Haase,<br />

W. Saak,A. Lützen,R. M. Gschwind,Org. Biomol. Chem.<br />

2005, 3,2691 – 2700.<br />

[203] Cleavage of covalent bonds,in the form of acid-mediated Nbutyloxycarbamate<br />

cleavage to reveal an ammonium binding<br />

site has been used to initiate degenerate shuttling,see: J. G.<br />

Cao,M. C. T. Fyfe,J. F. Stoddart,G. R. L. Cousins,P. T. Glink,<br />

J. Org. Chem. 2000, 65,1937 – 1946.<br />

[204] a) L. S. Jiang,J. Okano,A. Orita,J. Otera,Angew. Chem. 2004,<br />

116,2173 – 2176; Angew. Chem. Int. Ed. 2004, 43,2121 – 2124;<br />

b) for another system in which either metal coordination or<br />

organic substrate binding blocks the shuttling of the macrocycle<br />

in a [2]rotaxane,see D. A. Leigh,P. J. Lusby,A. M. Z. Slawin,<br />

D. B. Walker, Angew. Chem. 2005, 117,4633 – 4640; Angew.<br />

Chem. Int. Ed. 2005, 44,4557 – 4564.<br />

[205] D. A. Leigh,A. Troisi,F. Zerbetto,Angew. Chem. 2000, 112,<br />

358 – 361; Angew. Chem. Int. Ed. 2000, 39,350 – 353.<br />

[206] Even in 1991,when reporting the first degenerate molecular<br />

shuttle, [197] Stoddart noted that: “The opportunity now exists to<br />

desymmetrize the molecular shuttle by inserting nonidentical<br />

176 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

stations along the polyether thread in such a manner that<br />

these different stations can be addressed selectively by<br />

chemical,electrochemical,or photochemical means <strong>and</strong> so<br />

provide a mechanism to drive the bead to <strong>and</strong> fro between<br />

stations along the thread .”.<br />

[207] a) A. C. Benniston,A. Harriman,Angew. Chem. 1993, 105,<br />

1553 – 1555; Angew. Chem. Int. Ed. Engl. 1993, 32,1459 – 1461;<br />

b) A. C. Benniston,A. Harriman,V. M. Lynch,Tetrahedron<br />

Lett. 1994, 35,1473 – 1476; c) A. C. Benniston,A. Harriman,<br />

V. M. Lynch, J. Am. Chem. Soc. 1995, 117,5275 – 5291.<br />

[208] W. Clegg,C. Gimenez-Saiz,D. A. Leigh,A. Murphy,A. M. Z.<br />

Slawin,S. J. Teat,J. Am. Chem. Soc. 1999, 121,4124 – 4129.<br />

[209] M. Asakawa,G. Brancato,M. Fanti,D. A. Leigh,T. Shimizu,<br />

A. M. Z. Slawin,J. K. Y. Wong,F. Zerbetto,S. W. Zhang,J. Am.<br />

Chem. Soc. 2002, 124,2939 – 2950.<br />

[210] a) H. Murakami,A. Kawabuchi,K. Kotoo,M. Kunitake,N.<br />

Nakashima, J. Am. Chem. Soc. 1997, 119,7605 – 7606; b) H.<br />

Murakami,A. Kawabuchi,R. Matsumoto,T. Ido,N. Nakashima,<br />

J. Am. Chem. Soc. 2005, 127,15891 – 15899.<br />

[211] C. A. Stanier,S. J. Alderman,T. D. W. Claridge,H. L. Anderson,<br />

Angew. Chem. 2002, 114,1847 – 1850; Angew. Chem. Int.<br />

Ed. 2002, 41,1769 – 1772.<br />

[212] Intriguingly,insertion of an alkyl spacer beyond the viologen<br />

units restores the ability to undergo E!Z photoisomerization,<br />

concomitant with shuttling of the ring to this remote site (see<br />

Ref [210b]). Shuttling back to the central unit does not occur<br />

immediately on re-isomerization back to the E isomer,on<br />

account of the large kinetic barrier for passage over the<br />

viologen unit. It may be appropriate to regard this analogue,<br />

<strong>and</strong> therefore indeed 66,as two-station shuttles. However,any<br />

favorable interactions with the alkyl “stations” in either case<br />

are rather weak <strong>and</strong> poorly defined.<br />

[213] A molecular mechanics study of 66 yielded results consistent<br />

with the experimental observations for this shuttle. Although<br />

no comment is made of it,the optimized Z-66 co-conformation<br />

has the wider 3-rim of the a-CD closest to the azobenzene unit,<br />

in contrast to the experimental findings for 67,see: K. Sohlberg,<br />

B. G. Sumpter,D. W. Noid,J. Mol. Struct. 1999, 491,281 – 286.<br />

[214] For examples with cyclodextrin macrocycles,see Ref. [50].<br />

Similar effects have been observed when forming either<br />

pseudorotaxanes or rotaxanes by threading though the cavity<br />

in calixarenes,see: Ref. [181g] <strong>and</strong> A. Arduini,F. Ciesa,M.<br />

Fragassi,A. Pochini,A. Secchi,Angew. Chem. 2005, 117,282 –<br />

285; Angew. Chem. Int. Ed. 2005, 44,278 – 281.<br />

[215] Providing the chemical reaction is fast with respect to the<br />

movement of the macrocycle.<br />

[216] R. A. Bissell,E. Córdova,A. E. Kaifer,J. F. Stoddart,Nature<br />

1994, 369,133 – 137.<br />

[217] The subsequent proliferation of molecular machines based on<br />

CBPQT 4+ recognition has inspired a number of computational<br />

studies on complexation by this species,see: a) R. Castro,M. J.<br />

Berardi,E. Córdova,M. O. de Olza,A. E. Kaifer,J. D.<br />

Evanseck, J. Am. Chem. Soc. 1996, 118,10257 – 10268; b) R.<br />

Castro,K. R. Nixon,J. D. Evanseck,A. E. Kaifer,J. Org. Chem.<br />

1996, 61,7298 – 7303; c) R. Castro,P. D. Davidov,K. A. Kumar,<br />

A. P. March<strong>and</strong>,J. D. Evanseck,A. E. Kaifer,J. Phys. Org.<br />

Chem. 1997, 10,369 – 382; d) G. A. Kaminski,W. L. Jorgensen,<br />

J. Chem. Soc. Perkin Trans. 2 1999,2365 – 2375; e) A. T. Macias,<br />

K. A. Kumar,A. P. March<strong>and</strong>,J. D. Evanseck,J. Org. Chem.<br />

2000, 65,2083 – 2089; f) G. Ercolani,P. Mencarelli,J. Org.<br />

Chem. 2003, 68,6470 – 6473; g) K. C. Zhang,L. Liu,T. W. Mu,<br />

Q. X. Guo, Chem. Phys. Lett. 2001, 333,195 – 198; h) X.<br />

Grabuleda,P. Ivanov,C. Jaime,J. Org. Chem. 2003, 68,1539 –<br />

1547; these,in turn,have provoked the development of<br />

computational methods by which the related bistable shuttles<br />

themselves can be studied. Particular issues highlighted from<br />

this work are the importance of solvent interactions <strong>and</strong> thread<br />

Angew<strong>and</strong>te<br />

Chemie<br />

folding in determining the preferred co-conformation; see: i) X.<br />

Grabuleda,C. Jaime,J. Org. Chem. 1998, 63,9635 – 9643; j) X.<br />

Grabuleda,P. Ivanov,C. Jaime,J. Phys. Chem. B 2003, 107,<br />

7582 – 7588.<br />

[218] See: Ref. [179] <strong>and</strong> a) A. G. Kolchinski,D. H. Busch,N. W.<br />

Alcock, J. Chem. Soc. Chem. Commun. 1995,1289 – 1291;<br />

b) P. T. Glink,C. Schiavo,J. F. Stoddart,D. J. Williams,Chem.<br />

Commun. 1996,1483 – 1490; c) P. R. Ashton,P. T. Glink,J. F.<br />

Stoddart,P. A. Tasker,A. J. P. White,D. J. Williams,Chem. Eur.<br />

J. 1996, 2,729 – 736; d) S. J. Cantrill,A. R. Pease,J. F. Stoddart,<br />

J. Chem. Soc. Dalton Trans. 2000,3715 – 3734.<br />

[219] a) M.-V. Martínez-Díaz,N. Spencer,J. F. Stoddart,Angew.<br />

Chem. 1997, 109,1991 – 1994; Angew. Chem. Int. Ed. Engl.<br />

1997, 36,1904 – 1907; b) P. R. Ashton,R. Ballardini,V. Balzani,<br />

I. Baxter,A. Credi,M. C. T. Fyfe,M. T. G<strong>and</strong>olfi,M. Gómez-<br />

López,M.-V. Martínez-Díaz,A. Piersanti,N. Spencer,J. F.<br />

Stoddart,M. Venturi,A. J. P. White,D. J. Williams,J. Am.<br />

Chem. Soc. 1998, 120,11932 – 11942; c) S. GaraudØe,S. Silvi,<br />

M. Venturi,A. Credi,A. H. Flood,J. F. Stoddart,ChemPhys-<br />

Chem 2005, 6,2145 – 2152.<br />

[220] Terms such as “all-or-nothing” or “quantitative” are inappropriate<br />

descriptors for an equilibrium.<br />

[221] <strong>Molecular</strong> mechanics calculations on both forms of this system<br />

are in agreement with these experimental results,see: L.<br />

Frankfort,K. Sohlberg,J. Mol. Struct. 2003, 621,253 – 260.<br />

[222] For an example of another pH-switched shuttle from the<br />

Stoddart group,see: A. M. Elizarov,S.-H. Chiu,J. F. Stoddart,<br />

J. Org. Chem. 2002, 67,9175 – 9181.<br />

[223] a) J. D. Badjić,V. Balzani,A. Credi,S. Silvi,J. F. Stoddart,<br />

Science 2004, 303,1845 – 1849; b) J. D. Badjić,C. M. Ronconi,<br />

J. F. Stoddart,V. Balzani,S. Silvi,A. Credi,J. Am. Chem. Soc.<br />

2006, 128,1489 – 1499.<br />

[224] C. M. Keaveney,D. A. Leigh,Angew. Chem. 2004, 116,1242 –<br />

1244; Angew. Chem. Int. Ed. 2004, 43,1222 – 1224.<br />

[225] a) P. R. Ashton,R. A. Bissell,N. Spencer,J. F. Stoddart,M. S.<br />

Tolley, Synlett 1992,914 – 918; b) P. R. Ashton,R. A. Bissell,R.<br />

Gorski,D. Philp,N. Spencer,J. F. Stoddart,M. S. Tolley,Synlett<br />

1992,919 – 922; c) P. R. Ashton,R. A. Bissell,N. Spencer,J. F.<br />

Stoddart,M. S. Tolley,Synlett 1992,923 – 926; d) P. L. Anelli,<br />

M. Asakawa,P. R. Ashton,R. A. Bissell,G. Clavier,R. Gorski,<br />

A. E. Kaifer,S. J. Langford,G. Mattersteig,S. Menzer,D. Philp,<br />

A. M. Z. Slawin,N. Spencer,J. F. Stoddart,M. S. Tolley,D. J.<br />

Williams, Chem. Eur. J. 1997, 3,1113 – 1135; e) D. B. Amabilino,P.<br />

R. Ashton,S. E. Boyd,M. Gómez-López,W. Hayes,J. F.<br />

Stoddart, J. Org. Chem. 1997, 62,3062 – 3075.<br />

[226] Shuttle 69 also shows a certain degree of electrochemical<br />

switching in the deprotonated state when electrochemical<br />

reduction of the bipyridinium unit causes movement of the<br />

macrocycle back towards the amine moiety. In this state<br />

however,the electron-rich macrocycle has no significant<br />

favorable interactions with the thread <strong>and</strong> its position is not<br />

well-defined,see Ref. [219a].<br />

[227] H.-R. Tseng,S. A. Vignon,J. F. Stoddart,Angew. Chem. 2003,<br />

115,1529 – 1533; Angew. Chem. Int. Ed. 2003, 42,1491 – 1495.<br />

[228] a) M. Asakawa,P. R. Ashton,V. Balzani,A. Credi,C. Hamers,<br />

G. Mattersteig,M. Montalti,A. N. Shipway,N. Spencer,J. F.<br />

Stoddart,M. S. Tolley,M. Venturi,A. J. P. White,D. J. Williams,<br />

Angew. Chem. 1998, 110,357 – 361; Angew. Chem. Int. Ed. 1998,<br />

37,333 – 337; b) J. O. Jeppesen,J. Perkins,J. Becher,J. F.<br />

Stoddart, Angew. Chem. 2001, 113,1256 – 1261; Angew. Chem.<br />

Int. Ed. 2001, 40,1216 – 1221; c) J. O. Jeppesen,K. A. Nielsen,J.<br />

Perkins,S. A. Vignon,A. Di Fabio,R. Ballardini,M. T.<br />

G<strong>and</strong>olfi,M. Venturi,V. Balzani,J. Becher,J. F. Stoddart,<br />

Chem. Eur. J. 2003, 9,2982 – 3007; d) H.-R. Tseng,S. A.<br />

Vignon,P. C. Celestre,J. Perkins,J. O. Jeppesen,A. Di Fabio,<br />

R. Ballardini,M. T. G<strong>and</strong>olfi,M. Venturi,V. Balzani,J. F.<br />

Stoddart, Chem. Eur. J. 2004, 10,155 – 172; e) S. S. Kang,S. A.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

177


Reviews<br />

Vignon,H.-R. Tseng,J. F. Stoddart,Chem. Eur. J. 2004, 10,<br />

2555 – 2564; f) A. H. Flood,A. J. Peters,S. A. Vignon,D. W.<br />

Steuerman,H.-R. Tseng,S. Kang,J. R. Heath,J. F. Stoddart,<br />

Chem. Eur. J. 2004, 10,6558 – 6564; g) B. W. Laursen,S.<br />

Nygaard,J. O. Jeppesen,J. F. Stoddart,Org. Lett. 2004, 6,<br />

4167 – 4170; h) J. O. Jeppesen,S. Nygaard,S. A. Vignon,J. F.<br />

Stoddart, Eur. J. Org. Chem. 2005,196 – 220; i) J. W. Choi,A. H.<br />

Flood,D. W. Steuerman,S. Nygaard,A. B. Braunschweig,<br />

N. N. P. Moonen,B. W. Laursen,Y. Luo,E. DeIonno,A. J.<br />

Peters,J. O. Jeppesen,K. Xu,J. F. Stoddart,J. R. Heath,Chem.<br />

Eur. J. 2006, 12,261 – 279. For a review of the development of<br />

these systems,see Ref. [191ad].<br />

[229] a) A. M. Brouwer,C. Frochot,F. G. Gatti,D. A. Leigh,L.<br />

Mottier,F. Paolucci,S. Roffia,G. W. H. Wurpel,Science 2001,<br />

291,2124 – 2128; b) A. Altieri,F. G. Gatti,E. R. Kay,D. A.<br />

Leigh,D. Martel,F. Paolucci,A. M. Z. Slawin,J. K. Y. Wong,J.<br />

Am. Chem. Soc. 2003, 125,8644 – 8654.<br />

[230] For an independent computational study of this system which<br />

shows excellent agreement with the experimental results,see:<br />

X. G. Zheng,K. Sohlberg,J. Phys. Chem. A 2003, 107,1207 –<br />

1215.<br />

[231] N. Armaroli,V. Balzani,J.-P. Collin,P. Gaviæa,J.-P. Sauvage,B.<br />

Ventura, J. Am. Chem. Soc. 1999, 121,4397 – 4408.<br />

[232] For other examples of electrochemically switched shuttles,see:<br />

a) R. Ballardini,V. Balzani,W. Dehaen,A. E. Dell Erba,F. M.<br />

Raymo,J. F. Stoddart,M. Venturi,Eur. J. Org. Chem. 2000,<br />

591 – 602; b) P. R. Ashton,R. Ballardini,V. Balzani,A. Credi,<br />

K. R. Dress,E. Ishow,C. J. Kleverlaan,O. Kocian,J. A. Preece,<br />

N. Spencer,J. F. Stoddart,M. Venturi,S. Wenger,Chem. Eur. J.<br />

2000, 6,3558 – 3574 (for an example of light- <strong>and</strong> electrochemically<br />

triggered shuttling,see also Section 4.3.8); c) N. Kihara,<br />

M. Hashimoto,T. Takata,Org. Lett. 2004, 6,1693 – 1696.<br />

[233] a) S. A. Vignon,T. Jarrosson,T. Iijima,H.-R. Tseng,J. K. M.<br />

S<strong>and</strong>ers,J. F. Stoddart,J. Am. Chem. Soc. 2004, 126,9884 –<br />

9885; b) T. Iijima,S. A. Vignon,H.-R. Tseng,T. Jarrosson,<br />

J. K. M. S<strong>and</strong>ers,F. Marchioni,M. Venturi,E. Apostoli,V.<br />

Balzani,J. F. Stoddart,Chem. Eur. J. 2004, 10,6375 – 6392.<br />

[234] D. S. Marlin,D. Gonzµlez Cabrera,D. A. Leigh,A. M. Z.<br />

Slawin, Angew. Chem. 2006, 118,83 – 89; Angew. Chem. Int. Ed.<br />

2006, 45,77 – 83.<br />

[235] D. S. Marlin,D. Gonzµlez Cabrera,D. A. Leigh,A. M. Z.<br />

Slawin, Angew. Chem. 2006, 118,1413 – 1418; Angew. Chem.<br />

Int. Ed. 2006, 45,1385 – 1390.<br />

[236] The binding of Cd II ions to the BPA-glycine carbonyl oxygen<br />

atom lowers the pK a value of the adjacent NH proton,thus<br />

allowing selective deprotonation at this site. Model studies also<br />

show that deprotonation cannot occur while the macrocycle<br />

resides on the glycylglycine station <strong>and</strong> so the reaction must<br />

only occur in the minor translational isomer; once again,it is<br />

important to remember that the co-conformations in a simple<br />

[2]rotaxane are always in equilibrium.<br />

[237] D. A. Leigh,E. M. PØrez,Chem. Commun. 2004,2262 – 2263.<br />

[238] W. Abraham,L. Grubert,U. W. Grummt,K. Buck,Chem. Eur.<br />

J. 2004, 10,3562 – 3568.<br />

[239] S. Schmidt-Schäffer,L. Grubert,U. W. Grummt,K. Buck,W.<br />

Abraham, Eur. J. Org. Chem. 2006,378 – 398.<br />

[240] A. Altieri,G. Bottari,F. Dehez,D. A. Leigh,J. K. Y. Wong,F.<br />

Zerbetto, Angew. Chem. 2003, 115,2398 – 2402; Angew. Chem.<br />

Int. Ed. 2003, 42,2296 – 2300.<br />

[241] F. G. Gatti,D. A. Leigh,S. A. Nepogodiev,A. M. Z. Slawin,<br />

S. J. Teat,J. K. Y. Wong,J. Am. Chem. Soc. 2001, 123,5983 –<br />

5989.<br />

[242] V. Balzani,M. Clemente-León,A. Credi,B. Ferrer,M. Venturi,<br />

A. H. Flood,J. F. Stoddart,Proc. Natl. Acad. Sci. USA 2006,<br />

103,1178 – 1183. Similar to 73,shuttling in [2]rotaxane 81 can<br />

also be triggered by using an external source of electrons in an<br />

electrochemical cell.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

[243] G. W. H. Wurpel,A. M. Brouwer,I. H. M. van Stokkum,A.<br />

Farran,D. A. Leigh,J. Am. Chem. Soc. 2001, 123,11327 –<br />

11328.<br />

[244] Heat has been used to bring about cis!trans isomerization in<br />

80 that results in a concomitant net change of the position of the<br />

ring (see Ref. [240]),while a temperature increase has also<br />

been used to overcome a significant kinetic barrier to shuttling<br />

following a chemical change in a kinetically stable pseudorotaxane<br />

(see Ref. [188b]). In neither of these cases is the<br />

process reversible on cooling the material without some other<br />

stimulus being applied. In a number of the p-donor-/pacceptor-based<br />

molecular shuttles the population distribution<br />

can exhibit a certain degree of temperature sensitivity. This can<br />

occur when a large difference exists in the enthalpy of binding<br />

at each station <strong>and</strong>/or because of the formation of “alongside”<br />

interactions between the unoccupied electron-rich station <strong>and</strong><br />

the outer surface of the positively charged macrocycle,which<br />

incurs entropic losses <strong>and</strong> enthalpic gains. See,for example,<br />

Refs. [191ad, 228b,c,h,i].<br />

[245] G. Bottari,F. Dehez,D. A. Leigh,P. J. Nash,E. M. PØrez,<br />

J. K. Y. Wong,F. Zerbetto,Angew. Chem. 2003, 115,6066 –<br />

6069; Angew. Chem. Int. Ed. 2003, 42,5886 – 5889.<br />

[246] A C 12 chain has more than 500000 (3 12 ) possible C C rotamers<br />

<strong>and</strong> a significant number of these degrees of freedom must be<br />

lost upon forming the 83 structure.<br />

[247] A series of theoretical designs for entropy-driven mechanical<br />

motion in mechanically interlocked molecules have been<br />

suggested,see: A. Hanke,R. Metzler,Chem. Phys. Lett. 2002,<br />

359,22 – 26.<br />

[248] There are some clear limitations in using temperature as an<br />

entropy-driven switching mechanism: for example,it would be<br />

very difficult to address a small region of a larger array of<br />

shuttles in this way.<br />

[249] T. Da Ros,D. M. Guldi,A. F. Morales,D. A. Leigh,M. Prato,<br />

R. Turco, Org. Lett. 2003, 5,689 – 691.<br />

[250] a) C. G. Gong,H. W. Gibson,Angew. Chem. 1997, 109,2426 –<br />

2428; Angew. Chem. Int. Ed. Engl. 1997, 36,2331 – 2333;<br />

b) C. G. Gong,T. E. Glass,H. W. Gibson,Macromolecules<br />

1998, 31,308 – 313.<br />

[251] If it did,this would correspond to a type of information ratchet<br />

mechanism (see Section 1.4.2).<br />

[252] <strong>Molecular</strong>-scale logic systems which,in the main,do not rely on<br />

mechanical motions,<strong>and</strong> which operate through combinational<br />

logic processes,have been created. For reviews,see: a) P. Ball,<br />

Nature 2000, 406,118 – 120; b) A. P. de Silva,N. D. McClenaghan,C.<br />

P. McCoy in <strong>Molecular</strong> Switches (Ed.: B. L. Feringa),<br />

Wiley-VCH,Weinheim,2001,pp. 339 – 361; c) F. M. Raymo,<br />

Adv. Mater. 2002, 14,401 – 414; d) G. J. Brown,A. P. de Silva,S.<br />

Pagliari, Chem. Commun. 2002,2461 – 2463; e) V. Balzani,A.<br />

Credi,M. Venturi,ChemPhysChem 2003, 4,49 – 59; f) A. P.<br />

de Silva,N. D. McClenaghan,Chem. Eur. J. 2004, 10,574 – 586.<br />

[253] a) A. P. Malvino,J. A. Brown,Digital Computer Electronics,<br />

3rd ed.,Glencoe,Lake Forest,1993; b) R. J. Mitchell, Microprocessor<br />

Systems: An introduction,Macmillan,London,1995.<br />

[254] A series of stimuli-responsive molecular shuttles have been<br />

shown to exhibit relatively long-lived non-equilibrium (“metastable”)<br />

states when operating in self-assembled monolayers,a<br />

polymer electrolyte,or at low temperatures (see Refs. [191ad,<br />

228f,i] <strong>and</strong> Sections 8.1 <strong>and</strong> 8.3). This may account for the<br />

junction hysteresis observed in solid-state electronic devices<br />

that utilize such rotaxanes (see Section 8.3.1). Kinetically stable<br />

non-equilibrium co-conformations have also been observed in a<br />

cyclodextrin-based shuttle,see Ref. [210b].<br />

[255] The term “Brownian motor”,in the form of the German<br />

“Brownsche Motoren”,was introduced by Bartussek <strong>and</strong><br />

Hänggi (R. Bartussek,P. Hänggi,Phys. Bl. 1995, 51,506 –<br />

507) to describe load-bearing thermal ratchets <strong>and</strong> is also<br />

178 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

widely used in molecular biology. We have tried to place the<br />

term in a chemical context which is consistent with its meaning<br />

in biology <strong>and</strong> physics.<br />

[256] In principle,directional rotation of the macrocycle in a<br />

rotaxane could be induced by using rapid thermal fluctuations<br />

(or similar) <strong>and</strong> a chiral rotaxane (such as those in G. Brancato,<br />

F. Coutrot,D. A. Leigh,A. Murphy,J. K. Y. Wong,F. Zerbetto,<br />

Proc. Natl. Acad. Sci. USA 2002, 99,4967 – 4971)—see the<br />

“temperature ratchet” mechanism,Section 1.4.2. However,it<br />

cannot occur for a macrocycle undergoing Brownian motion in<br />

a chiral rotaxane at equilibrium,although this has been<br />

incorrectly suggested several times (O. Lukin,T. Kubota,Y.<br />

Okamoto,F. Schelhase,A. Yoneva,W. M. Müller,U. Müller,F.<br />

Vögtle, Angew. Chem. 2003, 115,4681 – 4684; Angew. Chem.<br />

Int. Ed. 2003, 42,4542 – 4545 <strong>and</strong> Ref. [191j]).<br />

[257] V. Bermudez,N. Capron,T. Gase,F. G. Gatti,F. Kajzar,D. A.<br />

Leigh,F. Zerbetto,S. W. Zhang,Nature 2000, 406,608 – 611.<br />

[258] a) A. Kapitulnik,S. Casalnuovo,K. C. Lim,A. J. Heeger,Phys.<br />

Rev. Lett. 1984, 53,469 – 472; b) K. C. Lim,A. Kapitulnik,R.<br />

Zacher,A. J. Heeger,J. Chem. Phys. 1985, 82,516 – 521.<br />

[259] F. G. Gatti,S. León,J. K. Y. Wong,G. Bottari,A. Altieri,<br />

M. A. F. Morales,S. J. Teat,C. Frochot,D. A. Leigh,A. M.<br />

Brouwer,F. Zerbetto,Proc. Natl. Acad. Sci. USA 2003, 100,10 –<br />

14.<br />

[260] a) R. Shukla,M. J. Deetz,B. D. Smith,Chem. Commun. 2000,<br />

2397 – 2398; b) M. J. Deetz,R. Shukla,B. D. Smith,Tetrahedron<br />

2002, 58,799 – 805; c) X. Fradera,M. Marquez,B. D. Smith,M.<br />

Orozco,F. J. Luque,J. Org. Chem. 2003, 68,4663 – 4673.<br />

[261] M. Linke,J.-C. Chambron,V. Heitz,J.-P. Sauvage,V. Semetey,<br />

Chem. Commun. 1998,2469 – 2470.<br />

[262] a) L. Raehm,J.-M. Kern,J.-P. Sauvage,Chem. Eur. J. 1999, 5,<br />

3310 – 3317; b) J.-M. Kern,L. Raehm,J.-P. Sauvage,B. Divisia-<br />

Blohorn,P. L. Vidal,Inorg. Chem. 2000, 39,1555 – 1560; c) N.<br />

Weber,C. Hamann,J.-M. Kern,J.-P. Sauvage,Inorg. Chem.<br />

2003, 42,6780 – 6792; d) I. Poleschak,J.-M. Kern,J.-P. Sauvage,<br />

Chem. Commun. 2004,474 – 476; e) U. LØtinois-Halbes,D.<br />

Hanss,J. M. Beierle,J.-P. Collin,J.-P. Sauvage,Org. Lett. 2005,<br />

7,5753 – 5756.<br />

[263] a) F. Vögtle,W. M. Müller,U. Müller,M. Bauer,K. Rissanen,<br />

Angew. Chem. 1993, 105,1356 – 1358; Angew. Chem. Int. Ed.<br />

Engl. 1993, 32,1295 – 1297; b) M. Bauer,W. M. Müller,U.<br />

Müller,K. Rissanen,F. Vögtle,Liebigs Ann. 1995,649 – 656.<br />

[264] P. Ceroni,D. A. Leigh,L. Mottier,F. Paolucci,S. Roffia,D.<br />

Tetard,F. Zerbetto,J. Phys. Chem. B 1999, 103,10171 – 10179.<br />

[265] M. J. Gunter,M. R. Johnston,J. Chem. Soc. Chem. Commun.<br />

1994,829 – 830.<br />

[266] L. Flamigni,A. M. Talarico,S. Serroni,F. Puntoriero,M. J.<br />

Gunter,M. R. Johnston,T. P. Jeynes,Chem. Eur. J. 2003, 9,<br />

2649 – 2659.<br />

[267] D. A. Leigh,K. Moody,J. P. Smart,K. J. Watson,A. M. Z.<br />

Slawin, Angew. Chem. 1996, 108,326 – 331; Angew. Chem. Int.<br />

Ed. Engl. 1996, 35,306 – 310.<br />

[268] Both experimental <strong>and</strong> computational studies have shown<br />

solvation effects to be important in determining the coconformational<br />

preference of certain cyclobis(paraquat-pphenylene)/diarene<br />

macrocyclic polyether [2]catenanes,see:<br />

a) P. R. Ashton,M. Blower,D. Philp,N. Spencer,J. F. Stoddart,<br />

M. S. Tolley,R. Ballardini,M. Ciano,V. Balzani,M. T.<br />

G<strong>and</strong>olfi,L. Prodi,C. H. McLean,New J. Chem. 1993, 17,<br />

689 – 695; b) F. M. Raymo,K. N. Houk,J. F. Stoddart,J. Org.<br />

Chem. 1998, 63,6523 – 6528; c) R. Ballardini,V. Balzani,M. T.<br />

G<strong>and</strong>olfi,R. E. Gillard,J. F. Stoddart,E. Tabellini,Chem. Eur.<br />

J. 1998, 4,449 – 459.<br />

[269] As well as demonstrating co-conformational preferences for<br />

both oxidation states which are in line with the experimental<br />

findings,a computational study of [2]catenane 92 also showed<br />

evidence for charge localization in the TTF unit of the singly<br />

Angew<strong>and</strong>te<br />

Chemie<br />

oxidized species,thus resulting in a barrierless,coulombic<br />

repulsion-driven TTF + !DNP shuttling step,see: X. G. Zheng,<br />

K. Sohlberg, Phys. Chem. Chem. Phys. 2004, 6,809 – 815; a<br />

further computational study employed a novel molecular<br />

dynamics approach to model the entire period of the shuttling<br />

motion (a timescale normally outside traditional molecular<br />

dynamics approaches). So-far only the TTF + !DNP step has<br />

been modeled,<strong>and</strong> important features to emerge include a<br />

virtually barrier-free potential-energy profile <strong>and</strong> a key role for<br />

counterions in the shuttling mechanism,see: M. Ceccarelli,F.<br />

Mercuri,D. Passerone,M. Parrinello,J. Phys. Chem. B 2005,<br />

109,17094 – 17099.<br />

[270] For other examples of switchable motion in [2]catenanes from<br />

the Stoddart group,see: a) P. R. Ashton,R. Ballardini,V.<br />

Balzani,M. T. G<strong>and</strong>olfi,D. J.-F. Marquis,L. PØrez-García,L.<br />

Prodi,J. F. Stoddart,M. Venturi,J. Chem. Soc. Chem. Commun.<br />

1994,177 – 180; b) P. R. Ashton,R. Ballardini,V. Balzani,A.<br />

Credi,M. T. G<strong>and</strong>olfi,S. Menzer,L. PØrez-García,L. Prodi,<br />

J. F. Stoddart,M. Venturi,A. J. P. White,D. J. Williams,J. Am.<br />

Chem. Soc. 1995, 117,11 171 – 11197 (redox switching); c) V.<br />

Balzani,A. Credi,S. J. Langford,F. M. Raymo,J. F. Stoddart,<br />

M. Venturi, J. Am. Chem. Soc. 2000, 122,3542 – 3543 (chemical<br />

switching); d) P. R. Ashton,V. Baldoni,V. Balzani,A. Credi,<br />

H. D. A. Hoffmann,M.-V. Martínez-Díaz,F. M. Raymo,J. F.<br />

Stoddart,M. Venturi,Chem. Eur. J. 2001, 7,3482 – 3493 (redox<br />

<strong>and</strong> pH switching).<br />

[271] Electrochemistry has been used to trigger co-conformational<br />

changes in a neutral p-donor/p-acceptor catenane; see: D. G.<br />

Hamilton,M. Montalti,L. Prodi,M. Fontani,P. Zanello,<br />

J. K. M. S<strong>and</strong>ers, Chem. Eur. J. 2000, 6,608 – 617.<br />

[272] B. Korybut-Daszkiewicz,A. Wie¸ckowska,R. Bilewicz,S.<br />

Domagała,K. Woz´niak, Angew. Chem. 2004, 116,1700 – 1704;<br />

Angew. Chem. Int. Ed. 2004, 43,1668 – 1672.<br />

[273] C. O. Dietrich-Buchecker,J.-P. Sauvage,J. P. Kintzinger,Tetrahedron<br />

Lett. 1983, 24,5095 – 5098.<br />

[274] a) C. O. Dietrich-Buchecker,J.-P. Sauvage,J.-M. Kern,J. Am.<br />

Chem. Soc. 1984, 106,3043 – 3045; b) M. Cesario,C. O.<br />

Dietrich-Buchecker,J. Guilhem,C. Pascard,J.-P. Sauvage,J.<br />

Chem. Soc. Chem. Commun. 1985,244 – 247; c) A.-M.<br />

Albrecht-Gary,Z. Saad,C. O. Dietrich-Buchecker,J.-P. Sauvage,<br />

J. Am. Chem. Soc. 1985, 107,3205 – 3209; d) C. Dietrich-<br />

Buchecker,J.-P. Sauvage,Tetrahedron 1990, 46,503 – 512.<br />

[275] a) M. Cesario,C. O. Dietrich,A. Edel,J. Guilhem,J. P.<br />

Kintzinger,C. Pascard,J.-P. Sauvage,J. Am. Chem. Soc. 1986,<br />

108,6250 – 6254; b) A.-M. Albrecht-Gary,C. Dietrich-<br />

Buchecker,Z. Saad,J.-P. Sauvage,J. Am. Chem. Soc. 1988,<br />

110,1467 – 1472; c) C. Dietrich-Buchecker,J.-P. Sauvage,J.-M.<br />

Kern, J. Am. Chem. Soc. 1989, 111,7791 – 7800; d) A.-M.<br />

Albrecht-Gary,C. Dietrich-Buchecker,Z. Saad,J.-P. Sauvage,<br />

J. Chem. Soc. Chem. Commun. 1992,280 – 282; e) N. Armaroli,<br />

L. Decola,V. Balzani,J.-P. Sauvage,C. O. Dietrich-Buchecker,<br />

J.-M. Kern,A. Bailal,J. Chem. Soc. Dalton Trans. 1993,3241 –<br />

3247.<br />

[276] Similar effects have been observed in a number of other metaltemplated<br />

interlocked systems,see: a) J.-P. Sauvage,J. Weiss,J.<br />

Am. Chem. Soc. 1985, 107,6108 – 6110; b) C. O. Dietrich-<br />

Buchecker,A. Khemiss,J.-P. Sauvage,J. Chem. Soc. Chem.<br />

Commun. 1986,1376 – 1378; c) B. Mohr,J.-P. Sauvage,R. H.<br />

Grubbs,M. Weck,Angew. Chem. 1997, 109,1365 – 1367;<br />

Angew. Chem. Int. Ed. Engl. 1997, 36,1308 – 1310; d) M.<br />

Koizumi,C. Dietrich-Buchecker,J.-P. Sauvage,Eur. J. Org.<br />

Chem. 2004,770 – 775.<br />

[277] D. B. Amabilino,C. O. Dietrich-Buchecker,A. Livoreil,L.<br />

PØrez-García,J.-P. Sauvage,J. F. Stoddart,J. Am. Chem. Soc.<br />

1996, 118,3905 – 3913.<br />

[278] A. Andrievsky,F. Ahuis,J. L. Sessler,F. Vögtle,D. Gudat,M.<br />

Moini, J. Am. Chem. Soc. 1998, 120,9712 – 9713.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

179


Reviews<br />

[279] A. M. L. Fuller,D. A. Leigh,P. J. Lusby,A. M. Z. Slawin,D. B.<br />

Walker, J. Am. Chem. Soc. 2005, 127,12612 – 12619.<br />

[280] D. A. Leigh,P. J. Lusby,A. M. Z. Slawin,D. B. Walker,Chem.<br />

Commun. 2005,4919 – 4921.<br />

[281] a) A. C. Laemmel,J.-P. Collin,J.-P. Sauvage,Eur. J. Inorg.<br />

Chem. 1999,383 – 386; b) E. Baranoff,J.-P. Collin,Y. Furusho,<br />

A. C. Laemmel,J.-P. Sauvage,Chem. Commun. 2000,1935 –<br />

1936; c) J.-P. Collin,A. C. Laemmel,J.-P. Sauvage,New J.<br />

Chem. 2001, 25,22 – 24; d) A. C. Laemmel,J.-P. Collin,J.-P.<br />

Sauvage,G. Accorsi,N. Armaroli,Eur. J. Inorg. Chem. 2003,<br />

467 – 474; e) S. Bonnet,J.-P. Collin,J.-P. Sauvage,Chem.<br />

Commun. 2005,3195 – 3197; f) J.-P. Collin,D. Jouvenot,M.<br />

Koizumi,J.-P. Sauvage,Inorg. Chem. 2005, 44,4693 – 4698.<br />

[282] P. Mobian,J.-M. Kern,J.-P. Sauvage,Angew. Chem. 2004, 116,<br />

2446 – 2449; Angew. Chem. Int. Ed. 2004, 43,2392 – 2395.<br />

[283] This reversible process has been shown to tolerate other<br />

catenane <strong>and</strong> also rotaxane structures,see: a) F. Arico,P.<br />

Mobian,J.-M. Kern,J.-P. Sauvage,Org. Lett. 2003, 5,1887 –<br />

1890; b) D. Pomeranc,D. Jouvenot,J.-C. Chambron,J.-P.<br />

Collin,V. Heitz,J.-P. Sauvage,Chem. Eur. J. 2003, 9,4247 –<br />

4254.<br />

[284] a) A. Livoreil,C. O. Dietrich-Buchecker,J.-P. Sauvage,J. Am.<br />

Chem. Soc. 1994, 116,9399 – 9400; b) F. Baumann,A. Livoreil,<br />

W. Kaim,J.-P. Sauvage,Chem. Commun. 1997,35 – 36; c) A.<br />

Livoreil,J.-P. Sauvage,N. Armaroli,V. Balzani,L. Flamigni,B.<br />

Ventura, J. Am. Chem. Soc. 1997, 119,12114 – 12 124.<br />

[285] D. J. Cardenas,A. Livoreil,J.-P. Sauvage,J. Am. Chem. Soc.<br />

1996, 118,11 980 – 11981.<br />

[286] D. A. Leigh,J. K. Y. Wong,F. Dehez,F. Zerbetto,Nature 2003,<br />

424,174 – 179.<br />

[287] A similar mechanism for achieving unidirectional rotation in a<br />

[2]catenane has been explored theoretically,see: R. D. Astumian,<br />

Proc. Natl. Acad. Sci. USA 2005, 102,1843 – 1847.<br />

[288] a) A. Troisi,M. A. Ratner,J. Am. Chem. Soc. 2002, 124,14528 –<br />

14529; b) P. E. Kornilovitch,A. M. Bratkovsky,R. S. Williams,<br />

Phys. Rev. B 2002, 66,245413; c) A. W. Ghosh,T. Rakshit,S.<br />

Datta, Nano Lett. 2004, 4,565 – 568.<br />

[289] A. Troisi,M. A. Ratner,Nano Lett. 2004, 4,591 – 595.<br />

[290] T. Seideman, J. Phys.: Condens. Matter 2003, 15,R521-R549.<br />

[291] a) K. Hoki,M. Yamaki,S. Koseki,Y. Fujimura,J. Chem. Phys.<br />

2003, 118,497 – 504; b) K. Hoki,M. Yamaki,Y. Fujimura,<br />

Angew. Chem. 2003, 115,3084 – 3086; Angew. Chem. Int. Ed.<br />

2003, 42,2976 – 2978.<br />

[292] In actual fact, 104 would be a poor choice practically as<br />

racemization is likely to occur through keto–enol tautomerism.<br />

However,this does not affect the validity of the principles<br />

involved.<br />

[293] K. Hoki,M. Yamaki,S. Koseki,Y. Fujimura,J. Chem. Phys.<br />

2003, 119,12 393 – 12398.<br />

[294] For theory,see: a) P. Reimann,M. Grifoni,P. Hänggi,Phys.<br />

Rev. Lett. 1997, 79,10 – 13; b) S. Yukawa,M. Kikuchi,G. Tatara,<br />

H. Matsukawa, J. Phys. Soc. Jpn. 1997, 66,2953 – 2956; c) I.<br />

Goychuk,P. Hänggi,Europhys. Lett. 1998, 43,503 – 509; for<br />

experiment,see Ref. [44p].<br />

[295] a) M. Yamaki,K. Hoki,Y. Ohtsuki,H. Kono,Y. Fujimura,J.<br />

Am. Chem. Soc. 2005, 127,7300 – 7301; b) M. Yamaki,K. Hoki,<br />

Y. Ohtsuki,H. Kono,Y. Fujimura,Phys. Chem. Chem. Phys.<br />

2005, 7,1900 – 1904.<br />

[296] For a review of the work in this area during the 1990s,see: A.<br />

Globus,C. W. Bauschlicher,J. Han,R. L. Jaffe,C. Levit,D.<br />

Srivastava, Nanotechnology 1998, 9,192 – 199.<br />

[297] a) K. E. Drexler, Engines of Creation: The Coming Era of<br />

Nanotechnology,Fourth Estate,London,1990; b) K. E. Drexler,<br />

Nanosystems: <strong>Molecular</strong> Machinery, Manufacturing <strong>and</strong><br />

Computation,Wiley,Chichester,1992.<br />

[298] a) J. C. Charlier,J. P. Michenaud,Phys. Rev. Lett. 1993, 70,<br />

1858 – 1861; b) R. C. Merkle, Nanotechnology 1993, 4,86 – 90;<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

c) Y. K. Kwon,D. Tomanek,Phys. Rev. B 1998, 58,R16001 –<br />

R16004; d) A. H. R. Palser, Phys. Chem. Chem. Phys. 1999, 1,<br />

4459 – 4464; e) A. N. Kolmogorov,V. H. Crespi,Phys. Rev. Lett.<br />

2000, 85,4727 – 4730; f) R. Saito,R. Matsuo,T. Kimura,G.<br />

Dresselhaus,M. S. Dresselhaus,Chem. Phys. Lett. 2001, 348,<br />

187 – 193; g) M. Damnjanovic,T. Vukovic,I. Milosevic,Eur.<br />

Phys. J. B 2002, 25,131 – 134; h) T. Vukovic,M. Damnjanovic,I.<br />

Milosevic, Physica E 2003, 16,259 – 268; i) M. Damnjanovic,E.<br />

Dobardzic,I. Milosevic,T. Vukovic,B. Nikolic,New J. Phys.<br />

2003, 6,148; j) J. Servantie,P. Gaspard,Phys. Rev. Lett. 2003,<br />

91,185503; k) A. V. Belikov,Y. E. Lozovik,A. G. Nikolaev,<br />

A. M. Popov, Chem. Phys. Lett. 2004, 385,72 – 78; l) E.<br />

Bichoutskaia,A. M. Popov,A. El-Barbary,M. I. Heggie,Y. E.<br />

Lozovik, Phys. Rev. B 2005, 71,113403.<br />

[299] a) R. E. Tuzun,D. W. Noid,B. G. Sumpter,Nanotechnology<br />

1995, 6,64 – 74; b) K. Sohlberg,R. E. Tuzun,B. G. Sumpter,<br />

D. W. Noid, Nanotechnology 1997, 8,103 – 111; c) R. E. Tuzun,<br />

K. Sohlberg,D. W. Noid,B. G. Sumpter,Nanotechnology 1998,<br />

9,37 – 48.<br />

[300] a) Y. E. Lozovik,A. Minogin,A. M. Popov,Phys. Lett. A 2003,<br />

313,112 – 121; b) Y. E. Lozovik,A. V. Minogin,A. M. Popov,<br />

JETP Lett. 2003, 77,631 – 635.<br />

[301] J. Han,A. Globus,R. Jaffe,G. Deardorff,Nanotechnology 1997,<br />

8,95 – 102.<br />

[302] R. E. Tuzun,D. W. Noid,B. G. Sumpter,Nanotechnology 1995,<br />

6,52 – 63.<br />

[303] D. Srivastava, Nanotechnology 1997, 8,186 – 192.<br />

[304] For an example of a proposed molecular bearing in which chiral<br />

nanotubes could exhibit unidirectional relative rotation driven<br />

by temperature variations (that is,a temperature ratchet),see:<br />

a) Z. C. Tu,Z. C. Ou-Yang,J. Phys.: Condens. Matter 2004, 16,<br />

1287 – 1292; for a similar case,driven by a fluctuating electric<br />

field (that is,a fluctuating potential ratchet),see: b) Z. C. Tu,X.<br />

Hu, Phys. Rev. B 2005, 72,033404.<br />

[305] a) L. Forró, Science 2000, 289,560 – 561; b) Q. S. Zheng,Q.<br />

Jiang, Phys. Rev. Lett. 2002, 88,045503; c) Q. S. Zheng,J. Z.<br />

Liu,Q. Jiang,Phys. Rev. B 2002, 65,245409; d) S. B. Legoas,<br />

V. R. Coluci,S. F. Braga,P. Z. Coura,S. O. Dantas,D. S.<br />

Galv¼o, Phys. Rev. Lett. 2003, 90,055504; e) W. Guo,Y. Guo,<br />

H. Gao,Q. Zheng,W. Zhong,Phys. Rev. Lett. 2003, 91,125501;<br />

f) Y. Zhao,C. C. Ma,G. H. Chen,Q. Jiang,Phys. Rev. Lett.<br />

2003, 91,175504; g) S. B. Legoas,V. R. Coluci,S. F. Braga,P. Z.<br />

Coura,S. Dantas,D. S. Galv¼o,Nanotechnology 2004, 15,<br />

S184 – S189; h) P. Tangney,S. G. Louie,M. L. Cohen,Phys. Rev.<br />

Lett. 2004, 93,065503.<br />

[306] J. W. Kang,H. J. Hwang,J. Appl. Phys. 2004, 96,3900 – 3905.<br />

[307] J. W. Kang,K. O. Song,O. K. Kwon,H. J. Hwang,Nanotechnology<br />

2005, 16,2670 – 2676.<br />

[308] D. Tasis,N. Tagmatarchis,A. Bianco,M. Prato,Chem. Rev.<br />

2006, 106,1105 – 1136.<br />

[309] a) J. Cumings,A. Zettl,Science 2000, 289,602 – 604; b) M. F.<br />

Yu,B. I. Yakobson,R. S. Ruoff,J. Phys. Chem. B 2000, 104,<br />

8764 – 8767.<br />

[310] The first examples of nanoelectromechanical systems incorporating<br />

moving nanotube components have been reported <strong>and</strong><br />

exemplify the state of the art in terms of both construction <strong>and</strong><br />

electric-field-driven manipulation for devices of this type,see:<br />

a) A. M. Fennimore,T. D. Yuzvinsky,W. Q. Han,M. S. Fuhrer,<br />

J. Cumings,A. Zettl,Nature 2003, 424,408 – 410; b) P. A.<br />

Williams,S. J. Papadakis,A. M. Patel,M. R. Falvo,S. Washburn,R.<br />

Superfine,Appl. Phys. Lett. 2003, 82,805 – 807; c) B.<br />

Bourlon,D. C. Glattli,C. Miko,L. Forró,A. Bachtold,Nano<br />

Lett. 2004, 4,709 – 712; d) S. J. Papadakis,A. R. Hall,P. A.<br />

Williams,L. Vicci,M. R. Falvo,R. Superfine,S. Washburn,<br />

Phys. Rev. Lett. 2004, 93,146101; e) B. C. Regan,S. Aloni,R. O.<br />

Ritchie,U. Dahmen,A. Zettl,Nature 2004, 428,924 – 927;<br />

f) B. C. Regan,S. Aloni,K. Jensen,R. O. Ritchie,A. Zettl,<br />

180 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Nano Lett. 2005, 5,1730 – 1733; g) J. C. Meyer,M. Paillet,S.<br />

Roth, Science 2005, 309,1539 – 1541; h) T. Ikuno,S.-I. Honda,T.<br />

Yasuda,K. Oura,M. Katayama,J. G. Lee,H. Mori,Appl. Phys.<br />

Lett. 2005, 87,213104; i) T. D. Yuzvinsky,A. M. Fennimore,A.<br />

Kis,A. Zettl,Nanotechnology 2006, 17,434 – 438.<br />

[311] In the computational studies of rotation in 104,Fujimura <strong>and</strong><br />

co-workers also considered motion driven by circularly polarised<br />

laser pulses. It was confirmed that the necessary asymmetry<br />

derives from the photon helicity,<strong>and</strong> rotational direction is<br />

independent of molecular chirality. The direction of rotation<br />

for a r<strong>and</strong>omly oriented population of rotors is therefore<br />

defined in the laboratory fixed frame (the frame of the applied<br />

field) <strong>and</strong> now the magnitude of rotation does depend on which<br />

enantiomer is used,see Ref. [291a].<br />

[312] J. Vacek,J. Michl,Proc. Natl. Acad. Sci. USA 2001, 98,5481 –<br />

5486.<br />

[313] For a review of such supramolecular grid structures,see: J.<br />

Michl,T. F. Magnera,Proc. Natl. Acad. Sci. USA 2002, 99,<br />

4788 – 4792.<br />

[314] For theoretical studies <strong>and</strong> preliminary experimental results on<br />

two structurally simpler azimuthal rotor (that is,axle perpendicular<br />

to the surface) systems,see: a) L. I. Clarke,D. Horinek,<br />

G. S. Kottas,N. Varaksa,T. F. Magnera,T. P. Hinderer,R. D.<br />

Horansky,J. Michl,J. C. Price,Nanotechnology 2002, 13,533 –<br />

540; b) D. Horinek,J. Michl,J. Am. Chem. Soc. 2003, 125,<br />

11900 – 11910.<br />

[315] a) X. L. Zheng,M. E. Mulcahy,D. Horinek,F. Galeotti,T. F.<br />

Magnera,J. Michl,J. Am. Chem. Soc. 2004, 126,4540 – 4542;<br />

b) T. F. Magnera,J. Michl,Top. Curr. Chem. 2005, 262,63 – 97.<br />

[316] At room temperature,it was observed that the molecules<br />

adopting a nonrotating conformation were interconverting with<br />

rotating forms on a timescale of minutes.<br />

[317] D. Horinek,J. Michl,Proc. Natl. Acad. Sci. USA 2005, 102,<br />

14175 – 14180.<br />

[318] At low temperatures (thermal energy less than maximum<br />

barrier heights),this motion is predicted to be near-synchronous<br />

given a strong enough electric field that is not too high in<br />

frequency. The mechanism is similar to a flashing ratchet except<br />

the electric field both varies the potential-energy surface <strong>and</strong><br />

powers motion over the surface. At temperatures which permit<br />

thermal crossing of the barrier,lower strength electric fields<br />

perform best,with the mechanism taking on the features of a<br />

true flashing Brownian ratchet. Interconversion between the<br />

residual diastereomers in the current structures is expected to<br />

scramble the unidirectional nature of the motion at ambient<br />

temperature. Analogues more resistant to isomerization<br />

should,however,be relatively easy to design.<br />

[319] a) D. C. Caskey,R. K. Shoemaker,J. Michl,Org. Lett. 2004, 6,<br />

2093 – 2096; b) D. C. Caskey,J. Michl,J. Org. Chem. 2005, 70,<br />

5442 – 5448.<br />

[320] H. Jian,J. M. Tour,J. Org. Chem. 2003, 68,5091 – 5103.<br />

[321] J. Karczmarek,J. Wright,P. Corkum,M. Ivanov,Phys. Rev. Lett.<br />

1999, 82,3420 – 3423.<br />

[322] D. M. Villeneuve,S. A. Aseyev,P. Dietrich,M. Spanner,M. Y.<br />

Ivanov,P. B. Corkum,Phys. Rev. Lett. 2000, 85,542 – 545.<br />

[323] An experimental example of direct momentum transfer from a<br />

linear flow to rotational motion of a chiral propeller-like<br />

molecule has been observed in a liquid-crystal monolayer (Y.<br />

Tabe,H. Yokoyama,Nat. Mater. 2003, 2,806 – 809). When a<br />

chiral liquid-crystal monolayer was deposited as a Langmuir<br />

film on a glycerol surface,a “target pattern” of exp<strong>and</strong>ing<br />

concentric circles was observed under reflected-light polarizing<br />

microscopy. The linear molecules align end-on but with a<br />

common tilt angle <strong>and</strong> undergo cooperative precession driven<br />

by the diffusion of water across the liquid/vapor interface.<br />

Molecules furthest from defects in the monolayer rotate most<br />

easily <strong>and</strong> transfer their changing orientation outwards,thereby<br />

Angew<strong>and</strong>te<br />

Chemie<br />

giving rise to the observed exp<strong>and</strong>ing pattern. The self-ordering<br />

properties of liquid crystals are essential for amplifying a small<br />

molecular-level effect which would otherwise be swamped by<br />

thermal motion.<br />

[324] J. Vacek,J. Michl,New J. Chem. 1997, 21,1259 – 1268.<br />

[325] Driving a nanoscale rotor constructed from carbon nanotube<br />

components with a flow of gas has recently been considered:<br />

J. W. Kang,H. J. Hwang,Nanotechnology 2004, 15,1633 – 1638.<br />

[326] K. Hoki,M. Sato,M. Yamaki,R. Sahnoun,L. Gonzµlez,S.<br />

Koseki,Y. Fujimura,J. Phys. Chem. B 2004, 108,4916 – 4921.<br />

[327] Y. Fujimura,L. Gonzµlez,D. Kröner,J. Manz,I. Mehdaoui,B.<br />

Schmidt, Chem. Phys. Lett. 2004, 386,248 – 253.<br />

[328] a) J. D. Dunitz,E. F. Maverick,K. N. Trueblood,Angew. Chem.<br />

1988, 100,910 – 926; Angew. Chem. Int. Ed. Engl. 1988, 27,880 –<br />

895; b) F. Horii,H. Kaji,H. Ishida,K. Kuwabara,K. Masuda,T.<br />

Tai, J. Mol. Struct. 1998, 441,303 – 311.<br />

[329] a) Z. Dominguez,H. Dang,M. J. Strouse,M. A. Garcia-<br />

Garibay, J. Am. Chem. Soc. 2002, 124,2398 – 2399; b) C. E.<br />

Godinez,G. Zepeda,M. A. Garcia-Garibay,J. Am. Chem. Soc.<br />

2002, 124,4701 – 4707; c) Z. Dominguez,H. Dang,M. J.<br />

Strouse,M. A. Garcia-Garibay,J. Am. Chem. Soc. 2002, 124,<br />

7719 – 7727; d) Z. Dominguez,T. A. V. Khuong,H. Dang,C. N.<br />

Sanrame,J. E. Nunez,M. A. Garcia-Garibay,J. Am. Chem.<br />

Soc. 2003, 125,8827 – 8837; e) T. A. V. Khuong,G. Zepeda,R.<br />

Ruiz,S. I. Khan,M. A. Garcia-Garibay,Cryst. Growth Des.<br />

2004, 4,15 – 18; f) C. E. Godinez,G. Zepeda,C. J. Mortko,H.<br />

Dang,M. A. Garcia-Garibay,J. Org. Chem. 2004, 69,1652 –<br />

1662; g) S. D. Karlen,M. A. Garcia-Garibay,Chem. Commun.<br />

2005,189 – 191; h) S. D. Karlen,S. I. Khan,M. A. Garcia-<br />

Garibay, Cryst. Growth Des. 2005, 5,53 – 55; i) S. D. Karlen,R.<br />

Ortiz,O. L. Chapman,M. A. Garcia-Garibay,J. Am. Chem.<br />

Soc. 2005, 127,6554 – 6555; j) R. D. Horansky,L. I. Clarke,J. C.<br />

Price,T. A. V. Khuong,P. D. Jarowski,M. A. Garcia-Garibay,<br />

Phys. Rev. B 2005, 72,014302.<br />

[330] For recent accounts of this work,see: a) M. A. Garcia-Garibay,<br />

Proc. Natl. Acad. Sci. USA 2005, 102,10 771 – 10776; b) S. D.<br />

Karlen,M. A. Garcia-Garibay,Top. Curr. Chem. 2005, 262,<br />

179 – 227.<br />

[331] An earlier,but related,structure is the “molecular turnstile”<br />

designed <strong>and</strong> created by Beddard <strong>and</strong> Moore: T. C. Bedard,<br />

J. S. Moore, J. Am. Chem. Soc. 1995, 117,10662 – 10 671.<br />

[332] The diborate congener of 110,in which the trityl carbon atoms<br />

are replaced by boron,has also been reported: J. R. Gardinier,<br />

P. J. Pellechia,M. D. Smith,J. Am. Chem. Soc. 2005, 127,<br />

12448 – 12449. Compared to 110,a more open packing<br />

arrangement is found in the crystalline state owing to coulombic<br />

repulsion which results in a lower barrier to rotary motion.<br />

[333] T. Shima,F. Hampel,J. A. Gladysz,Angew. Chem. 2004, 116,<br />

5653 – 5656; Angew. Chem. Int. Ed. 2004, 43,5537 – 5540.<br />

[334] T. Akutagawa,K. Shitagami,S. Nishihara,S. Takeda,T.<br />

Hasegawa,T. Nakamura,Y. Hosokoshi,K. Inoue,S. Ikeuchi,<br />

Y. Miyazaki,K. Saito,J. Am. Chem. Soc. 2005, 127,4397 – 4402.<br />

[335] For an overview of recent experimental results in nanoscale<br />

locomotion,see: G. A. Ozin,I. Manners,S. Fournier-Bidoz,A.<br />

Arsenault, Adv. Mater. 2005, 17,3011 – 3018.<br />

[336] a) N. O. Young,J. S. Goldstein,M. J. Block,J. Fluid Mech. 1959,<br />

6,350 – 356; b) A. W. Adamson,Physical Chemistry of Surfaces,<br />

5th ed.,Wiley,New York,1990.<br />

[337] For an historical perspective,see: L. E. Scriven,C. V. Sternling,<br />

Nature 1960, 187,186 – 188.<br />

[338] a) K. D. Barton,R. S. Subramanian,J. Colloid Interface Sci.<br />

1989, 133,211 – 222; b) A. M. Cazabat,F. Heslot,S. M. Troian,<br />

P. Carles, Nature 1990, 346,824 – 826.<br />

[339] a) M. A. Burns,C. H. Mastrangelo,T. S. Sammarco,F. P. Man,<br />

J. R. Webster,B. N. Johnson,B. Foerster,D. Jones,Y. Fields,<br />

A. R. Kaiser,D. T. Burke,Proc. Natl. Acad. Sci. USA 1996, 93,<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

181


Reviews<br />

5556 – 5561; b) D. E. Kataoka,S. M. Troian,Nature 1999, 402,<br />

794 – 797.<br />

[340] a) C. D. Bain,G. D. Burnetthall,R. R. Montgomerie,Nature<br />

1994, 372,414 – 415; b) D. F. Dossantos,T. Ondarçuhu,Phys.<br />

Rev. Lett. 1995, 75,2972 – 2975; c) M. E. R. Shanahan,P. G.<br />

de Gennes, C. R. Acad. Sci. Ser. IIb 1997, 324,261 – 268; d) P. G.<br />

de Gennes, Europhys. Lett. 1997, 39,407 – 412; e) P. G. de Gennes,<br />

Physica A 1998, 249,196 – 205; f) S. W. Lee,P. E.<br />

Laibinis, J. Am. Chem. Soc. 2000, 122,5395 – 5396; g) Y.<br />

Sumino,N. Magome,T. Hamada,K. Yoshikawa,Phys. Rev.<br />

Lett. 2005, 94,068301.<br />

[341] A similar effect can be induced in mercury droplets by a redox<br />

reaction. This may be controlled by application of an electric<br />

field in suitable electrolytes <strong>and</strong> has been used to move mercury<br />

droplets by a Brownian ratchet mechanism,see Refs. [44e,g].<br />

Another related phenomenon is observed because of the<br />

solvent-swelling of amphiphilic polymer gels: T. Mitsumata,K.<br />

Ikeda,J. P. Gong,Y. Osada,Appl. Phys. Lett. 1998, 73,2366 –<br />

2368.<br />

[342] R. F. Ismagilov,A. Schwartz,N. Bowden,G. M. Whitesides,<br />

Angew. Chem. 2002, 114,674 – 676; Angew. Chem. Int. Ed.<br />

2002, 41,652 – 654.<br />

[343] W. F. Paxton,K. C. Kistler,C. C. Olmeda,A. Sen,S. K. St<br />

Angelo,Y. Y. Cao,T. E. Mallouk,P. E. Lammert,V. H. Crespi,<br />

J. Am. Chem. Soc. 2004, 126,13424 – 13431.<br />

[344] a) R. Golestanian,T. B. Liverpool,A. Ajdari,Phys. Rev. Lett.<br />

2005, 94,220801; b) W. E. Paxton,A. Sen,T. E. Mallouk,<br />

Chem. Eur. J. 2005, 11,6462 – 6470; an alternative mechanism<br />

considered herein is “self-electrophoresis”,whereby two ends<br />

of a conducting object catalyze two different electrochemical<br />

half-cell reactions,essentially generating its own electic field.<br />

This would result in a flow of electrons within the particle,<br />

accompanied by a concomitant balancing flow of protons (or<br />

other electrolyte ions) across its surface,thus resulting in<br />

propulsion of the object with respect to the fluid. Experimentally,this<br />

mechanism has very recently been investigated as a<br />

method for mixing a solution over a metallic surface,see:<br />

c) T. R. Kline,W. F. Paxton,Y. Wang,D. Velegol,T. E.<br />

Mallouk,A. Sen,J. Am. Chem. Soc. 2005, 127,17150 – 17151;<br />

<strong>and</strong> in an example where enzymes are empolyed as the redoxactive<br />

species to propel a macroscopic conducting carbon fiber,<br />

see: d) N. Mano,A. Heller,J. Am. Chem. Soc. 2005, 127,<br />

11574 – 11575.<br />

[345] P. Dhar,T. M. Fischer,Y. Wang,T. E. Mallouk,W. F. Paxton,A.<br />

Sen, Nano Lett. 2006, 6,66 – 72.<br />

[346] J. M. Catchmark,S. Subramanian,A. Sen,Small 2005, 1,202 –<br />

206.<br />

[347] T. R. Kline,W. F. Paxton,T. E. Mallouk,A. Sen,Angew. Chem.<br />

2005, 117,754 – 756; Angew. Chem. Int. Ed. 2005, 44,744 – 746.<br />

[348] S. Fournier-Bidoz,A. C. Arsenault,I. Manners,G. A. Ozin,<br />

Chem. Commun. 2005,441 – 443.<br />

[349] J. Vicario,R. Eelkema,W. R. Browne,A. Meetsma,R. M. La<br />

Crois,B. L. Feringa,Chem. Commun. 2005,3936 – 3938.<br />

[350] G. Taylor, Proc. R. Soc. London Ser. A 1951, 209,447 – 461.<br />

[351] a) L. E. Becker,S. A. Koehler,H. A. Stone,J. Fluid Mech.<br />

2003, 490,15 – 35; b) R. Dreyfus,J. Baudry,H. A. Stone,Eur.<br />

Phys. J. B 2005, 47,161 – 164.<br />

[352] a) A. Najafi,R. Golestanian,Phys. Rev. E 2004, 69,062901;<br />

b) A. Najafi,R. Golestanian,J. Phys.: Condens. Matter 2005, 17,<br />

S1203 – S1208.<br />

[353] a) A. Shapere,F. Wilczek,Phys. Rev. Lett. 1987, 58,2051 – 2054;<br />

b) A. Shapere,F. Wilczek,J. Fluid Mech. 1989, 198,557 – 585;<br />

c) A. Shapere,F. Wilczek,J. Fluid Mech. 1989, 198,587 – 599;<br />

d) H. A. Stone,A. D. T. Samuel,Phys. Rev. Lett. 1996, 77,<br />

4102 – 4104; e) E. M. Purcell, Proc. Natl. Acad. Sci. USA 1997,<br />

94,11307 – 11311; f) C. H. Wiggins,R. E. Goldstein,Phys. Rev.<br />

Lett. 1998, 80,3879 – 3882; g) A. Ajdari,H. A. Stone,Phys.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

Fluids 1999, 11,1275 – 1277; h) S. Camalet,F. Jülicher,J. Prost,<br />

Phys. Rev. Lett. 1999, 82,1590 – 1593; i) J. E. Avron,O. Gat,O.<br />

Kenneth, Phys. Rev. Lett. 2004, 93,186001; j) J. E. Avron,O.<br />

Kenneth,D. H. Oaknin,New J. Phys. 2005, 7,234; two related<br />

mechanisms for mechanical non-Brownian propulsion across a<br />

periodic but symmetrical surface have also recently been<br />

considered. These can be regarded as “self-propelled” walkers<br />

(see Section 4.4) in which all steps in the mechanism involve<br />

changes to the moving object,while the track remains passive<br />

throughout,see: k) M. Porto,M. Urbakh,J. Klafter,Phys. Rev.<br />

Lett. 2000, 84,6058 – 6061; l) Z. Wang,Phys. Rev. E 2004, 70,<br />

031903.<br />

[354] R. Dreyfus,J. Baudry,M. L. Roper,M. Fermigier,H. A. Stone,<br />

J. Bibette, Nature 2005, 437,862 – 865.<br />

[355] D.-H. Qu,Q.-C. Wang,H. Tian,Angew. Chem. 2005, 117,5430 –<br />

5433; Angew. Chem. Int. Ed. 2005, 44,5296 – 5299.<br />

[356] The suggestion of photo/thermal imine isomerization was made<br />

to us by Jean-Marie Lehn in 2004 during discussions on possible<br />

means to break time-reversal symmetry in chemical reactions.<br />

These ideas were recently published in the context of rotary<br />

motors: J.-M. Lehn, Chem. Eur. J. 2006, 12,5910 – 5915.<br />

[357] a) N. B. Baranova,B. Y. Zel dovich,Chem. Phys. Lett. 1978, 57,<br />

435 – 437; b) B. Space,H. Rabitz,A. Lörincz,P. Moore,J. Chem.<br />

Phys. 1996, 105,9515 – 9524.<br />

[358] a) G. Binnig,H. Rohrer,C. Gerber,E. Weibel,Phys. Rev. Lett.<br />

1982, 49,57 – 61; b) G. Binnig,H. Rohrer,Rev. Mod. Phys. 1987,<br />

59,615 – 625.<br />

[359] G. Binnig,C. F. Quate,C. Gerber,Phys. Rev. Lett. 1986, 56,<br />

930 – 933.<br />

[360] For general reviews on the use of probe microscopies to<br />

manipulate single atoms <strong>and</strong> molecules,see: a) special issue on<br />

“Force <strong>and</strong> Tunneling Microscopy”: Chem. Rev. 1997, 97,1015 –<br />

1230; b) J. K. Gimzewski,C. Joachim,Science 1999, 283,1683 –<br />

1688; for reviews on the early development of single atom <strong>and</strong><br />

molecule STM manipulation,see: c) J. A. Stroscio,D. M.<br />

Eigler, Science 1991, 254,1319 – 1326; d) P. Avouris,Acc.<br />

Chem. Res. 1995, 28,95 – 102; for reviews on the development<br />

of STM manipulation of large molecules,see: e) J. K. Gimzewski,T.<br />

A. Jung,M. T. Cuberes,R. R. Schlittler,Surf. Sci. 1997,<br />

386,101 – 114; f) F. Rosei,M. Schunack,Y. Naitoh,P. Jiang,A.<br />

Gourdon,E. Laegsgaard,I. Stensgaard,C. Joachim,F. Besenbacher,<br />

Prog. Surf. Sci. 2003, 71,95 – 146; for reviews on the<br />

development of <strong>and</strong> advances in AFM,see: g) H. Takano,J. R.<br />

Kenseth,S. S. Wong,J. C. O Brien,M. D. Porter,Chem. Rev.<br />

1999, 99,2845 – 2890; h) M. Guthold,M. Falvo,W. G. Matthews,S.<br />

Paulson,J. Mullin,S. Lord,D. Erie,S. Washburn,R.<br />

Superfine,F. P. Brooks,R. M. Taylor,J. Mol. Graphics Modell.<br />

1999, 17,187 – 197; i) F. J. Giessibl,Rev. Mod. Phys. 2003, 75,<br />

949 – 983; for reviews on the use of AFM to probe the<br />

intramolecular interactions in single biomolecules <strong>and</strong> intermolecular<br />

interactions of biological complexes,see: j) C.<br />

Bustamante,D. Keller,Phys. Today 1995, 48(12),32 – 38;<br />

k) A. Engel,H. E. Gaub,D. J. Müller,Curr. Biol. 1999, 9,<br />

R133 – R136; l) H. Clausen-Schaumann,M. Seitz,R. Krautbauer,H.<br />

E. Gaub,Curr. Opin. Chem. Biol. 2000, 4,524 – 530;<br />

m) B. Samori, Chem. Eur. J. 2000, 6,4249 – 4255; n) A. Janshoff,<br />

M. Neitzert,Y. Oberdörfer,H. Fuchs,Angew. Chem. 2000, 112,<br />

3346 – 3374; Angew. Chem. Int. Ed. 2000, 39,3212 – 3237; o) M.<br />

Rief,H. Grubmüller,ChemPhysChem 2002, 3,255 – 261;<br />

p) R. B. Best,J. Clarke,Chem. Commun. 2002,183 – 192; for<br />

reviews on the use of STM to induce single-molecule chemical<br />

reactions,see: q) S. W. Hla,G. Meyer,K.-H. Rieder,Chem-<br />

PhysChem 2001, 2,361 – 366; for reviews on the role of probe<br />

microscopies in the development of molecular electronics,see:<br />

Ref. [290] <strong>and</strong> r) C. Joachim,J. K. Gimzewski,Proc. IEEE<br />

1998, 86,184 – 190; s) C. Joachim,J. K. Gimzewski,A. Aviram,<br />

182 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Nature 2000, 408,541 – 548; t) F. Moresco,Phys. Rep. 2004, 399,<br />

175 – 225.<br />

[361] For a futuristic vision of molecular manufacturing based on the<br />

atom-by-atom manipulation of matter by nanoscale machines,<br />

see Ref. [297].<br />

[362] D. M. Eigler,C. P. Lutz,W. E. Rudge,Nature 1991, 352,600 –<br />

603.<br />

[363] T. A. Jung,R. R. Schlittler,J. K. Gimzewski,H. Tang,C.<br />

Joachim, Science 1996, 271,181 – 184.<br />

[364] Free rotation around the porphyrin–phenyl bond means that<br />

different leg–porphyrin angles are observed on different<br />

surfaces (from virtually perpendicular to coplanar). In all<br />

cases,however,the planar porphyrin core is insulated from the<br />

surface by the bulky tert-butyl groups: a) T. A. Jung,R. R.<br />

Schlittler,J. K. Gimzewski,Nature 1997, 386,696 – 698; b) F.<br />

Moresco,G. Meyer,K.-H. Rieder,H. Ping,H. Tang,C.<br />

Joachim, Surf. Sci. 2002, 499,94 – 102.<br />

[365] Note this method of “pushing” the molecule away from a<br />

moving STM tip is in fact different from the strategy used to<br />

move single atoms <strong>and</strong> simple molecules at low temperature,<br />

where the mobile species is trapped directly under the tip by<br />

electrostatic or van der Waals interactions before “sliding” to<br />

the new position.<br />

[366] A modified technique was required to effect motion of the<br />

same molecule at low temperature; however this then allowed a<br />

detailed description of the precise conformational changes <strong>and</strong><br />

their effect on the tunneling current between the tip <strong>and</strong><br />

substrate,see: a) F. Moresco,G. Meyer,K.-H. Rieder,H. Tang,<br />

A. Gourdon,C. Joachim,Appl. Phys. Lett. 2001, 78,306 – 308;<br />

b) F. Moresco,G. Meyer,K.-H. Rieder,H. Tang,A. Gourdon,<br />

C. Joachim, Phys. Rev. Lett. 2001, 87,088302.<br />

[367] F. Biscarini,W. Gebauer,D. Di Domenico,R. Zamboni,J. I.<br />

Pascual,D. A. Leigh,A. Murphy,D. Tetard,Synth. Met. 1999,<br />

102,1466 – 1467.<br />

[368] P. Samori,F. Jackel,O. Unsal,A. Godt,J. P. Rabe,ChemPhys-<br />

Chem 2001, 2,461 – 464.<br />

[369] H. Shigekawa,K. Miyake,J. Sumaoka,A. Harada,M.<br />

Komiyama, J. Am. Chem. Soc. 2000, 122,5411 – 5412.<br />

[370] M. T. Cuberes,R. R. Schlittler,J. K. Gimzewski,Appl. Phys.<br />

Lett. 1996, 69,3016 – 3018.<br />

[371] M. Feng,X. Guo,X. Lin,X. He,W. Ji,S. Du,D. Zhang,D. Zhu,<br />

H. Gao, J. Am. Chem. Soc. 2005, 127,15338 – 15339.<br />

[372] K. Kitagawa,T. Morita,S. Kimura,Angew. Chem. 2005, 117,<br />

6488 – 6491; Angew. Chem. Int. Ed. 2005, 44,6330 – 6333.<br />

[373] A related phenomenon has been reported for gold-supported<br />

self-assembled monolayers of a rotaxane. A change in surface<br />

height was observed after applying a positive voltage with a<br />

conducting AFM tip,see: D. Wu,X. Yin,X. Zhang,H.-R.<br />

Tseng,J. F. Stoddart,IEEE Trans: Nanotechnol. 2003, 2,606 –<br />

608.<br />

[374] J. K. Gimzewski,C. Joachim,R. R. Schlittler,V. Langlais,H.<br />

Tang,I. Johannsen,Science 1998, 281,531 – 533.<br />

[375] For an example of an STM-imaged rotation of a guest molecule<br />

in a cavity within a 2D lattice,where the guest <strong>and</strong> lattice<br />

molecules are different,see: S. J. H. Griessl,M. Lackinger,F.<br />

Jamitzky,T. Markert,M. Hietschold,W. M. Heckl,Langmuir<br />

2004, 20,9403 – 9407.<br />

[376] For an interesting discussion which places the aforementioned<br />

results in the context of ultimately creating,<strong>and</strong> observing,a<br />

single-molecule unidirectional rotor,see: C. Joachim,J. K.<br />

Gimzewski, Struct. Bonding (Berlin) 2001, 99,1 – 18.<br />

[377] The incorporation of a “small-molecule” fragment into a<br />

polymer of known force–extension behavior can be used to<br />

circumvent such problems in certain cases.<br />

[378] a) M. Cavallini,F. Biscarini,S. León,F. Zerbetto,G. Bottari,<br />

D. A. Leigh, Science 2003, 299,531 – 531; b) J. F. Moulin,J. C.<br />

Kengne,R. Kshirsagar,M. Cavallini,F. Biscarini,S. León,F.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Zerbetto,G. Bottari,D. A. Leigh,J. Am. Chem. Soc. 2006, 128,<br />

526 – 532.<br />

[379] a) F. Moresco,G. Meyer,K.-H. Rieder,H. Tang,A. Gourdon,<br />

C. Joachim, Phys. Rev. Lett. 2001, 86,672 – 675; b) C. Loppacher,M.<br />

Guggisberg,O. Pfeiffer,E. Meyer,M. Bammerlin,<br />

R. Luthi,R. Schlittler,J. K. Gimzewski,H. Tang,C. Joachim,<br />

Phys. Rev. Lett. 2003, 90,066107.<br />

[380] C. Viala,A. Seechi,A. Gourdon,Eur. J. Org. Chem. 2002,<br />

4185 – 4189.<br />

[381] a) A. Gourdon, Eur. J. Org. Chem. 1998,2797 – 2801; b) V. J.<br />

Langlais,R. R. Schlittler,H. Tang,A. Gourdon,C. Joachim,<br />

J. K. Gimzewski, Phys. Rev. Lett. 1999, 83,2809 – 2812; c) F.<br />

Moresco,L. Gross,L. Grill,M. Alemani,A. Gourdon,C.<br />

Joachim,K.-H. Rieder,Appl. Phys. A 2005, 80,913 – 920,<strong>and</strong><br />

references therein; this work has recently been summarized,<br />

see: d) F. Moresco,A. Gourdon,Proc. Natl. Acad. Sci. USA<br />

2005, 102,8809 – 8814.<br />

[382] L. Soukiassian,A. J. Mayne,G. Comtet,L. Hellner,G.<br />

Dujardin,A. Gourdon,J. Chem. Phys. 2005, 122,134704.<br />

[383] M. Lastapis,M. Martin,D. Riedel,L. Hellner,G. Comtet,G.<br />

Dujardin, Science 2005, 308,1000 – 1003.<br />

[384] a) C. Joachim,H. Tang,F. Moresco,G. Rapenne,G. Meyer,<br />

Nanotechnology 2002, 13,330 – 335; b) G. Jimenez-Bueno,G.<br />

Rapenne, Tetrahedron Lett. 2003, 44,6261 – 6263.<br />

[385] G. Rapenne, Org. Biomol. Chem. 2005, 3,1165 – 1169.<br />

[386] L. Grill,K.-H. Rieder,F. Moresco,G. Jimenez-Bueno,C. Wang,<br />

G. Rapenne,C. Joachim,Surf. Sci. 2005, 584,L153 L158.<br />

[387] Nanoscale friction is still a matter of intense investigation <strong>and</strong><br />

debate,see: a) B. Bhushan,J. N. Israelachvili,U. L<strong>and</strong>man,<br />

Nature 1995, 374,607 – 616; b) B. Luan,M. O. Robbins,Nature<br />

2005, 435,929 – 932.<br />

[388] a) A. Buldum,J. P. Lu,Phys. Rev. Lett. 1999, 83,5050 – 5053;<br />

b) B. Ni,S. B. Sinnott,Surf. Sci. 2001, 487,87 – 96; c) J. W. Kang,<br />

H. J. Hwang, Nanotechnology 2004, 15,614 – 621; d) N. Sasaki,<br />

K. Miura, Jpn. J. Appl. Phys. 2004, 43,4486 – 4491.<br />

[389] a) M. R. Falvo,R. M. Taylor,A. Helser,V. Chi,F. P. Brooks,S.<br />

Washburn,R. Superfine,Nature 1999, 397,236 – 238; b) M. R.<br />

Falvo,J. Steele,R. M. Taylor,R. Superfine,Phys. Rev. B 2000,<br />

62,R10665 – R10667; c) K. Miura,T. Takagi,S. Kamiya,T.<br />

Sahashi,M. Yamauchi,Nano Lett. 2001, 1,161 – 163.<br />

[390] a) P. Moriarty,Y. R. Ma,M. D. Upward,P. H. Beton,Surf. Sci.<br />

1998, 407,27 – 35; b) K. Miura,S. Kamiya,N. Sasaki,Phys. Rev.<br />

Lett. 2003, 90,055509; c) D. L. Keeling,M. J. Humphry,R. H. J.<br />

Fawcett,P. H. Beton,C. Hobbs,L. Kantorovich,Phys. Rev. Lett.<br />

2005, 94,146104.<br />

[391] Y. Shirai,A. J. Osgood,Y. Zhao,K. F. Kelly,J. M. Tour,Nano<br />

Lett. 2005, 5,2330 – 2334.<br />

[392] a) B. Das,K. L. Sebastian,Chem. Phys. Lett. 2000, 330,433 –<br />

439; b) B. Das,K. L. Sebastian,Chem. Phys. Lett. 2002, 357,<br />

25 – 31.<br />

[393] Y. Fujikawa,J. T. Sadowski,K. F. Kelly,K. S. Nakayama,E. T.<br />

Mickelson,R. H. Hauge,J. L. Margrave,T. Sakurai,Jpn. J.<br />

Appl. Phys. 2002, 41,245 – 249.<br />

[394] a) S. J. Stranick,M. M. Kamna,P. S. Weiss,Science 1994, 266,<br />

99 – 102; b) R. Otero,F. Hummelink,F. Sato,S. B. Legoas,P.<br />

Thostrup,E. Laegsgaard,I. Stensgaard,D. S. Galv¼o,F.<br />

Besenbacher, Nat. Mater. 2004, 3,779 – 782.<br />

[395] K. Y. Kwon,K. L. Wong,G. Pawin,L. Bartels,S. Stolbov,T. S.<br />

Rahman, Phys. Rev. Lett. 2005, 95,166101.<br />

[396] A. Carella,G. Rapenne,J. P. Launay,New J. Chem. 2005, 29,<br />

288 – 290.<br />

[397] A. Carella,J. Jaud,G. Rapenne,J. P. Launay,Chem. Commun.<br />

2003,2434 – 2435.<br />

[398] For reviews,see: a) S. M. Block,Nature 1992, 360,493 – 495;<br />

b) S. C. Kuo,M. P. Sheetz,Trends Cell Biol. 1992, 2,116 – 118;<br />

c) T. Yanagida,Y. Harada,A. Ishijima,Trends Biochem. Sci.<br />

1993, 18,319 – 324; d) K. Svoboda,S. M. Block,Annu. Rev.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

183


Reviews<br />

Biophys. Biomol. Struct. 1994, 23,247 – 285; e) A. D. Mehta,M.<br />

Rief,J. A. Spudich,D. A. Smith,R. M. Simmons,Science 1999,<br />

283,1689 – 1695; f) C. Bustamante,J. C. Macosko,G. J. L.<br />

Wuite, Nat. Rev. Mol. Cell Biol. 2000, 1,130 – 136; g) J. F.<br />

Allem<strong>and</strong>,D. Bensimon,V. Croquette,Curr. Opin. Struct. Biol.<br />

2003, 13,266 – 274; h) X. Zhuang,Science 2004, 305,188 – 190;<br />

i) K. C. Neuman,S. M. Block,Rev. Sci. Instrum. 2004, 75,2787 –<br />

2809.<br />

[399] Y. Arai,R. Yasuda,K. Akashi,Y. Harada,H. Miyata,K.<br />

Kinosita,H. Itoh,Nature 1999, 399,446 – 448.<br />

[400] a) P. Y. Chiou,A. T. Ohta,M. C. Wu,Nature 2005, 436,370 –<br />

372; b) K. Dholakia, Nat. Mater. 2005, 4,579 – 580.<br />

[401] For reviews,see: a) W. E. Moerner,T. BaschØ,Angew. Chem.<br />

1993, 105,537 – 557; Angew. Chem. Int. Ed. Engl. 1993, 32,457 –<br />

476; b) W. E. Moerner, Science 1994, 265,46 – 53; c) W. E.<br />

Moerner, Acc. Chem. Res. 1996, 29,563 – 571; d) X. S. Xie,Acc.<br />

Chem. Res. 1996, 29,598 – 606; e) P. M. Goodwin,W. P.<br />

Ambrose,R. A. Keller,Acc. Chem. Res. 1996, 29,607 – 613;<br />

f) S. M. Nie,R. N. Zare,Annu. Rev. Biophys. Biomol. Struct.<br />

1997, 26,567 – 596; g) T. Plakhotnik,E. A. Donley,U. P. Wild,<br />

Annu. Rev. Phys. Chem. 1997, 48,181 – 212; h) X. S. Xie,J. K.<br />

Trautman, Annu. Rev. Phys. Chem. 1998, 49,441 – 480; i) W. E.<br />

Moerner,M. Orrit,Science 1999, 283,1670 – 1676; j) S. Weiss,<br />

Science 1999, 283,1676 – 1683; k) W. P. Ambrose,P. M. Goodwin,J.<br />

H. Jett,A. Van Orden,J. H. Werner,R. A. Keller,Chem.<br />

Rev. 1999, 99,2929 – 2956; l) A. A. Deniz,T. A. Laurence,M.<br />

Dahan,D. S. Chemla,P. G. Schultz,S. Weiss,Annu. Rev. Phys.<br />

Chem. 2001, 52,233 – 253; m) W. E. Moerner,J. Phys. Chem. B<br />

2002, 106,910 – 927; n) R. A. Keller,W. P. Ambrose,A. A.<br />

Arias,H. Gai,S. R. Emory,P. M. Goodwin,J. H. Jett,Anal.<br />

Chem. 2002, 74,316a – 324a; o) special issue on “Single-<br />

Molecule Physics <strong>and</strong> Chemistry”: J. Chem. Phys. 2002, 117,<br />

10923–11033; p) M. Bohmer,J. Enderlein,ChemPhysChem<br />

2003, 4,793 – 808; q) W. E. Moerner,D. P. Fromm,Rev. Sci.<br />

Instrum. 2003, 74,3597 – 3619; r) P. Tinnefeld,M. Sauer,Angew.<br />

Chem. 2005, 117,2698 – 2728; Angew. Chem. Int. Ed. 2005, 44,<br />

2642 – 2671; s) special issue on “Single-Molecule Spectroscopy”:<br />

Acc. Chem. Res. 2005, 38,503 – 610.<br />

[402] N. Fang,H. Lee,C. Sun,X. Zhang,Science 2005, 308,534 – 537.<br />

[403] For reviews see: a) R. M. Crooks,A. J. Ricco,Acc. Chem. Res.<br />

1998, 31,219 – 227; b) L. M. Goldenberg,M. R. Bryce,M. C.<br />

Petty, J. Mater. Chem. 1999, 9,1957 – 1974; c) S. Flink,<br />

F. C. J. M. van Veggel,D. N. Reinhoudt,Adv. Mater. 2000, 12,<br />

1315 – 1328; d) A. N. Shipway,E. Katz,I. Willner,ChemPhys-<br />

Chem 2000, 1,18 – 52; e) E. Katz,I. Willner,Electroanalysis<br />

2003, 15,913 – 947; f) G. Cooke,Angew. Chem. 2003, 115,5008 –<br />

5018; Angew. Chem. Int. Ed. 2003, 42,4860 – 4870; g) D. Astruc,<br />

M.-C. Daniel,J. Ruiz,Chem. Commun. 2004,2637 – 2649; h) U.<br />

Drechsler,B. Erdogan,V. M. Rotello,Chem. Eur. J. 2004, 10,<br />

5570 – 5579.<br />

[404] A reversible recognition event has also been observed in a<br />

nanoporous silicate matrix,see: a) S. Y. Chia,J. Cao,J. F.<br />

Stoddart,J. I. Zink,Angew. Chem. 2001, 113,2513 – 2517;<br />

Angew. Chem. Int. Ed. 2001, 40,2447 – 2451; for other<br />

molecular-recognition processes occurring within solids,see:<br />

b) T. E. Mallouk,J. A. Gavin,Acc. Chem. Res. 1998, 31,209 –<br />

217.<br />

[405] R. Hern<strong>and</strong>ez,H.-R. Tseng,J. W. Wong,J. F. Stoddart,J. I.<br />

Zink, J. Am. Chem. Soc. 2004, 126,3370 – 3371.<br />

[406] T. Gase,D. Gr<strong>and</strong>o,P. A. Chollet,F. Kajzar,A. Murphy,D. A.<br />

Leigh, Adv. Mater. 1999, 11,1303 – 1305.<br />

[407] For a computational study of the conformations <strong>and</strong> coconformational<br />

dynamics of the same molecule on a graphite<br />

surface,see: M. S. Deleuze,J. Am. Chem. Soc. 2000, 122,1130 –<br />

1143.<br />

[408] G. Bidan,M. Billon,B. Divisia-Blohorn,J.-M. Kern,L. Raehm,<br />

J.-P. Sauvage, New J. Chem. 1998, 22,1139 – 1141.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

[409] L. Raehm,J.-M. Kern,J.-P. Sauvage,C. Hamann,S. Palacin,J. P.<br />

Bourgoin, Chem. Eur. J. 2002, 8,2153 – 2162.<br />

[410] H.-R. Tseng,D. M. Wu,N. X. L. Fang,X. Zhang,J. F. Stoddart,<br />

ChemPhysChem 2004, 5,111 – 116.<br />

[411] T. J. Huang,H.-R. Tseng,L. Sha,W. X. Lu,B. Brough,A. H.<br />

Flood,B.-D. Yu,P. C. Celestre,J. P. Chang,J. F. Stoddart,C.-M.<br />

Ho, Nano Lett. 2004, 4,2065 – 2071.<br />

[412] B. Long,K. Nikitin,D. Fitzmaurice,J. Am. Chem. Soc. 2003,<br />

125,15490 – 15498.<br />

[413] K. Nikitin,D. Fitzmaurice,J. Am. Chem. Soc. 2005, 127,8067 –<br />

8076.<br />

[414] R. A. van Delden,M. K. J. ter Wiel,M. M. Pollard,J. Vicario,<br />

N. Koumura,B. L. Feringa,Nature 2005, 437,1337 – 1340.<br />

[415] Attachment of triptycene-based rotors (such as those discussed<br />

in Sections 2.1.1 <strong>and</strong> 2.1.2) to surfaces has been simulated<br />

computationally,see: S. M. Hou,T. Sagara,D. C. Xu,T. R.<br />

Kelly,E. Ganz,Nanotechnology 2003, 14,566 – 570.<br />

[416] a) T. E. Glass, J. Am. Chem. Soc. 2000, 122,4522 – 4523; b) J.<br />

Raker,T. E. Glass,Tetrahedron 2001, 57,10 233 – 10240; c) J.<br />

Raker,T. E. Glass,J. Org. Chem. 2001, 66,6505 – 6512; d) J.<br />

Raker,T. E. Glass,J. Org. Chem. 2002, 67,6113 – 6116.<br />

[417] For an example of binding-induced conformational control<br />

leading to changes in the electronic properties of oligothiophenes,see:<br />

a) B. Jousselme,P. Blanchard,E. Levillain,J.<br />

Delaunay,M. Allain,P. Richomme,D. Rondeau,N. Gallego-<br />

Planas,J. Roncali,J. Am. Chem. Soc. 2003, 125,1363 – 1370; in<br />

polyether-substituted poly(thiophene)s—<strong>and</strong> in other conjugated<br />

polymers—changes in electronic properties can often be<br />

attributed to electrostatic interactions on ion doping. For<br />

examples where binding-induced conformational changes of<br />

the polymer backbone are thought to play a role,see: b) J.<br />

Roncali,R. Garreau,D. Delabouglise,F. Garnier,M. Lemaire,<br />

J. Chem. Soc. Chem. Commun. 1989,679 – 681; c) J. Roncali,<br />

L. H. Shi,F. Garnier,J. Phys. Chem. 1991, 95,8983 – 8989;<br />

d) M. J. Marsella,T. M. Swager,J. Am. Chem. Soc. 1993, 115,<br />

12214 – 12215; e) M. J. Marsella,R. J. Newl<strong>and</strong>,P. J. Carroll,<br />

T. M. Swager, J. Am. Chem. Soc. 1995, 117,9842 – 9848; for an<br />

example of similar effects in bipyridyl-substitued poly(phenylvinylene),see:<br />

f) B. Wang,M. R. Wasielewski,J. Am. Chem.<br />

Soc. 1997, 119,12 – 21.<br />

[418] For reviews on functionalised poly(thiophene)s,see: a) J.<br />

Roncali, J. Mater. Chem. 1999, 9,1875 – 1893; b) D. T.<br />

McQuade,A. E. Pullen,T. M. Swager,Chem. Rev. 2000, 100,<br />

2537 – 2574; c) G. Barbarella,M. Melucci,G. Sotgiu,Adv.<br />

Mater. 2005, 17,1581 – 1593.<br />

[419] Similar changes to electronic properties can be observed on<br />

application of a photonic stimulus to oligothiophenes appended<br />

with an azobenzene moiety,see: a) B. Jousselme,P. Blanchard,<br />

N. Gallego-Planas,J. Delaunay,M. Allain,P. Richomme,E.<br />

Levillain,J. Roncali,J. Am. Chem. Soc. 2003, 125,2888 – 2889;<br />

b) B. Jousselme,P. Blanchard,N. Gallego-Planas,E. Levillain,<br />

J. Delaunay,M. Allain,P. Richomme,J. Roncali,Chem. Eur. J.<br />

2003, 9,5297 – 5306.<br />

[420] D. H. R. Barton, Science 1970, 169,539 – 544.<br />

[421] J. Clayden,A. Lund,L. S. Vallverdu,M. Helliwell,Nature 2004,<br />

431,966 – 971.<br />

[422] a) H. Yuasa,N. Miyagawa,T. Izumi,M. Nakatani,M. Izumi,H.<br />

Hashimoto, Org. Lett. 2004, 6,1489 – 1492; for other examples<br />

of chair-flip-based conformational switches,see: b) C. Monahan,J.<br />

T. Bien,B. D. Smith,Chem. Commun. 1998,431 – 432.<br />

[423] M. Barboiu,L. Prodi,M. Montalti,N. Zaccheroni,N. Kyritsakas,J.-M.<br />

Lehn,Chem. Eur. J. 2004, 10,2953 – 2959.<br />

[424] M. Linke-Schaetzel,C. E. Anson,A. K. Powell,G. Buth,E.<br />

Palomares,J. D. Durrant,T. S. Balaban,J.-M. Lehn,Chem. Eur.<br />

J. 2006, 12,1931 – 1940; for an example of using the cationtriggered<br />

conformational change of terpyridines to vary the<br />

through-space electronic properties of two quinone substitu-<br />

184 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

ents,see: M. Buschel,M. Helldobler,J. Daub,Chem. Commun.<br />

2002,1338 – 1339.<br />

[425] a) J. Berninger,R. Krauss,H. G. Weinig,U. Koert,B. Ziemer,<br />

K. Harms, Eur. J. Org. Chem. 1999,875 – 884; b) R. Krauss,<br />

H. G. Weinig,M. Seydack,J. Bendig,U. Koert,Angew. Chem.<br />

2000, 112,1905 – 1908; Angew. Chem. Int. Ed. 2000, 39,1835 –<br />

1837; c) H. G. Weinig,R. Krauss,M. Seydack,J. Bendig,U.<br />

Koert, Chem. Eur. J. 2001, 7,2075 – 2088.<br />

[426] U. Koert,R. Krauss,H. G. Weinig,C. Heumann,B. Ziemer,C.<br />

Mugge,M. Seydack,J. Bendig,Eur. J. Org. Chem. 2001,575 –<br />

586.<br />

[427] M. Karle,D. Bockelmann,D. Schumann,C. Griesinger,U.<br />

Koert, Angew. Chem. 2003, 115,4684 – 4687; Angew. Chem. Int.<br />

Ed. 2003, 42,4546 – 4549.<br />

[428] V. A. Azov,A. Schlegel,F. Diederich,Angew. Chem. 2005, 117,<br />

4711 – 4715; Angew. Chem. Int. Ed. 2005, 44,4635 – 4638.<br />

[429] a) M. C. JimØnez,C. Dietrich-Buchecker,J.-P. Sauvage,Angew.<br />

Chem. 2000, 112,3422 – 3425; Angew. Chem. Int. Ed. 2000, 39,<br />

3284 – 3287; b) M. C. JimØnez-Molero,C. Dietrich-Buchecker,<br />

J.-P. Sauvage, Chem. Eur. J. 2002, 8,1456 – 1466.<br />

[430] E. M. PØrez,D. T. F. Dryden,D. A. Leigh,G. Teobaldi,F.<br />

Zerbetto, J. Am. Chem. Soc. 2004, 126,12 210 – 12211.<br />

[431] G. Bottari,D. A. Leigh,E. M. PØrez,J. Am. Chem. Soc. 2003,<br />

125,13360 – 13361.<br />

[432] In a related photoswitched system,the position of a pyridiniumcontaining<br />

macrocycle is used to influence the interaction<br />

between two fluorophores attached to the thread,thus illiciting<br />

an optical response on shuttling,see a) Y. J. Li,H. Li,Y. L. Li,<br />

H. B. Liu,S. Wang,X. R. He,N. Wang,D. B. Zhu,Org. Lett.<br />

2005, 7,4835 – 4838; modification of a FRET response between<br />

fluorophores appended to the macrocycle <strong>and</strong> thread has also<br />

been demonstrated in a solvent-switchable hydrogen-bonded<br />

system,see b) H. Onagi,J. Rebek,Jr.,Chem. Commun. 2005,<br />

4604 – 4606.<br />

[433] Q.-C. Wang,D.-H. Qu,J. Ren,K. Chen,H. Tian,Angew. Chem.<br />

2004, 116,2715 – 2719; Angew. Chem. Int. Ed. 2004, 43,2661 –<br />

2665.<br />

[434] During synthesis,the directional isomer of (E)-133·2H shown<br />

in Scheme 79 was formed exclusively but seems to arise by a<br />

kinetic trapping mechanism rather than from the thermodynamic<br />

stability of this product,see: Section 4.3.1 <strong>and</strong> Ref. [50k].<br />

[435] a) D.-H. Qu,Q.-C. Wang,J. Ren,H. Tian,Org. Lett. 2004, 6,<br />

2085 – 2088; b) D.-H. Qu,Q.-C. Wang,H. Tian,Mol. Cryst. Liq.<br />

Cryst. 2005, 430,59 – 65.<br />

[436] As well as a photochemical switching mechanism,it was found<br />

in one example that the position of the ring—<strong>and</strong> therefore the<br />

fluorescence intensities—could also be controlled by temperature,see<br />

Ref. [435b].<br />

[437] D.-H. Qu,Q.-C. Wang,X. Ma,H. Tian,Chem. Eur. J. 2005, 11,<br />

5929 – 5937.<br />

[438] J. Elemans,E. J. A. Bijsterveld,A. E. Rowan,R. J. M. Nolte,<br />

Chem. Commun. 2000,2443 – 2444.<br />

[439] a) P. Thordarson,E. J. A. Bijsterveld,A. E. Rowan,R. J. M.<br />

Nolte, Nature 2003, 424,915 – 918; b) P. Thordarson,R. J. M.<br />

Nolte,A. E. Rowan,Aust. J. Chem. 2004, 57,323 – 327; c) P. R.<br />

Carlier, Angew. Chem. 2004, 116,2654 – 2657; Angew. Chem.<br />

Int. Ed. 2004, 43,2602 – 2605.<br />

[440] For detailed accounts of the development of these devices,see<br />

Refs. [191h,x,y].<br />

[441] a) C. P. Collier,E. W. Wong,M. Belohradsky´,F. M. Raymo,J. F.<br />

Stoddart,P. J. Kuekes,R. S. Williams,J. R. Heath,Science 1999,<br />

285,391 – 394; b) E. W. Wong,C. P. Collier,M. Bìhloradsky´,<br />

F. M. Raymo,J. F. Stoddart,J. R. Heath,J. Am. Chem. Soc.<br />

2000, 122,5831 – 5840.<br />

[442] a) R. C. Ahuja,P.-L. Caruso,D. Möbius,G. Wildburg,H.<br />

Ringsdorf,D. Philp,J. A. Preece,J. F. Stoddart,Langmuir 1993,<br />

9,1534 – 1544; b) R. C. Ahuja,P.-L. Caruso,D. Möbius,D.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

Philp,J. A. Preece,H. Ringsdorf,J. F. Stoddart,G. Wildburg,<br />

Thin Solid Films 1996, 285,671 – 677; c) C. L. Brown,U. Jonas,<br />

J. A. Preece,H. Ringsdorf,M. Seitz,J. F. Stoddart,Langmuir<br />

2000, 16,1924 – 1930; d) M. Asakawa,M. Higuchi,G. Mattersteig,T.<br />

Nakamura,A. R. Pease,F. M. Raymo,T. Shimizu,J. F.<br />

Stoddart, Adv. Mater. 2000, 12,1099 – 1102.<br />

[443] C. P. Collier,G. Mattersteig,E. W. Wong,Y. Luo,K. Beverly,J.<br />

Sampaio,F. M. Raymo,J. F. Stoddart,J. R. Heath,Science 2000,<br />

289,1172 – 1175.<br />

[444] C. P. Collier,J. O. Jeppesen,Y. Luo,J. Perkins,E. W. Wong,<br />

J. R. Heath,J. F. Stoddart,J. Am. Chem. Soc. 2001, 123,12632 –<br />

12641.<br />

[445] Y. Luo,C. P. Collier,J. O. Jeppesen,K. A. Nielsen,E. Delonno,<br />

G. Ho,J. Perkins,H.-R. Tseng,T. Yamamoto,J. F. Stoddart,<br />

J. R. Heath, ChemPhysChem 2002, 3,519 – 525.<br />

[446] D. W. Steuerman,H.-R. Tseng,A. J. Peters,A. H. Flood,J. O.<br />

Jeppesen,K. A. Nielsen,J. F. Stoddart,J. R. Heath,Angew.<br />

Chem. 2004, 116,6648 – 6653; Angew. Chem. Int. Ed. 2004, 43,<br />

6486 – 6491.<br />

[447] W.-Q. Deng,R. P. Muller,W. A. Goddard,J. Am. Chem. Soc.<br />

2004, 126,13 562 – 13563.<br />

[448] a) I. C. Lee,C. W. Frank,T. Yamamoto,H.-R. Tseng,A. H.<br />

Flood,J. F. Stoddart,Langmuir 2004, 20,5809 – 5828; b) K.<br />

Nørgaard,J. O. Jeppesen,B. A. Laursen,J. B. Simonsen,M. J.<br />

Weyg<strong>and</strong>,K. Kjaer,J. F. Stoddart,T. Bjørnholm,J. Phys. Chem.<br />

B 2005, 109,1063 – 1066.<br />

[449] S. S. Jang,Y. H. Jang,Y. H. Kim,W. A. Goddard,J. W. Choi,<br />

J. R. Heath,B. W. Laursen,A. H. Flood,J. F. Stoddart,K.<br />

Nørgaard,T. Bjørnholm,J. Am. Chem. Soc. 2005, 127,14804 –<br />

14816.<br />

[450] Y. H. Jang,S. G. Hwang,Y. H. Kim,S. S. Jang,W. A. Goddard,<br />

J. Am. Chem. Soc. 2004, 126,12636 – 12645.<br />

[451] a) S. S. Jang,Y. H. Jang,Y. H. Kim,W. A. Goddard,A. H.<br />

Flood,B. W. Laursen,H.-R. Tseng,J. F. Stoddart,J. O. Jeppesen,J.<br />

W. Choi,D. W. Steuerman,E. Delonno,J. R. Heath,J.<br />

Am. Chem. Soc. 2005, 127,1563 – 1575; b) Y. H. Jang,S. S. Jang,<br />

W. A. Goddard, J. Am. Chem. Soc. 2005, 127,4959 – 4964.<br />

[452] Y. H. Kim,S. S. Jang,Y. H. Jang,W. A. Goddard,Phys. Rev.<br />

Lett. 2005, 94,156801.<br />

[453] J. H. Huang,A. H. Flood,C.-W. Chu,S. Kang,T.-F. Guo,T.<br />

Yamamoto,H.-R. Tseng,B.-D. Yu,Y. Yang,J. F. Stoddart,C.-<br />

M. Ho, IEEE Trans: Nanotechnol. 2003, 2,698 – 701.<br />

[454] K. Nørgaard,B. W. Laursen,S. Nygaard,K. Kjaer,H.-R. Tseng,<br />

A. H. Flood,J. F. Stoddart,T. Bjørnholm,Angew. Chem. 2005,<br />

117,7197 – 7201; Angew. Chem. Int. Ed. 2005, 44,7035 – 7039.<br />

[455] a) H. B. Yu,Y. Luo,K. Beverly,J. F. Stoddart,H.-R. Tseng,<br />

J. R. Heath, Angew. Chem. 2003, 115,5884 – 5889; Angew.<br />

Chem. Int. Ed. 2003, 42,5706 – 5711; b) Y. Chen,D. A. A.<br />

Ohlberg,X. M. Li,D. R. Stewart,R. S. Williams,J. O. Jeppesen,<br />

K. A. Nielsen,J. F. Stoddart,D. L. Olynick,E. Anderson,Appl.<br />

Phys. Lett. 2003, 82,1610 – 1612; c) Y. Chen,G. Y. Jung,<br />

D. A. A. Ohlberg,X. M. Li,D. R. Stewart,J. O. Jeppesen,<br />

K. A. Nielsen,J. F. Stoddart,R. S. Williams,Nanotechnology<br />

2003, 14,462 – 468; d) D. R. Stewart,D. A. A. Ohlberg,P. A.<br />

Beck,Y. Chen,R. S. Williams,J. O. Jeppesen,K. A. Nielsen,<br />

J. F. Stoddart, Nano Lett. 2004, 4,133 – 136.<br />

[456] J. R. Heath,M. A. Ratner,Phys. Today 2003, 56(5),43 – 49. In<br />

fact,Williams <strong>and</strong> co-workers have investigated the possibility<br />

of using filament formation between two metal electrodes,<br />

separated by an insulating organic single-molecule layer,to<br />

create a switching device,see: C. N. Lau,D. R. Stewart,R. S.<br />

Williams,M. Bockrath,Nano Lett. 2004, 4,569 – 572.<br />

[457] M. R. Diehl,D. W. Steuerman,H.-R. Tseng,S. A. Vignon,A.<br />

Star,P. C. Celestre,J. F. Stoddart,J. R. Heath,ChemPhysChem<br />

2003, 4,1335 – 1339.<br />

[458] a) K. Ichimura, Chem. Rev. 2000, 100,1847 – 1873; b) N.<br />

Tamaoki, Adv. Mater. 2001, 13,1135 – 1147; c) T. Ikeda,A.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

185


Reviews<br />

Kanazawa in <strong>Molecular</strong> Switches (Ed.: B. L. Feringa),Wiley-<br />

VCH,Weinheim,2001,pp. 363 – 397.<br />

[459] a) T. Kosa,E. Weinan,P. Palffy-Muhoray,Int. J. Eng. Sci. 2000,<br />

38,1077 – 1084; b) P. Palffy-Muhoray,T. Kosa,E. Weinan,Mol.<br />

Cryst. Liq. Cryst. 2002, 375,577 – 591; c) P. Palffy-Muhoray,T.<br />

Kosa,E. Weinan,Appl. Phys. A 2002, 75,293 – 300; d) P. Palffy-<br />

Muhoray,T. Kosa,W. E,Braz. J. Phys. 2002, 32,552 – 563.<br />

[460] R. A. van Delden,T. Mecca,C. Rosini,B. L. Feringa,Chem.<br />

Eur. J. 2004, 10,61 – 70.<br />

[461] a) For the switching of the cholesteric screw sense by photoconversion<br />

between two pseudoenantiomeric overcrowded<br />

alkene diastereomers,see: B. L. Feringa,N. P. M. Huck,H. A.<br />

V<strong>and</strong>oren, J. Am. Chem. Soc. 1995, 117,9929 – 9930; b) for<br />

inversion of h<strong>and</strong>edness in the liquid-crystal phase by generating<br />

an enantiomeric excess (of extremely low magnitude) in an<br />

overcrowded alkene dopant by irradiation with circularly<br />

polarized light,see: N. P. M. Huck,W. F. Jager,B. de Lange,<br />

B. L. Feringa, Science 1996, 273,1686 – 1688; c) for reversal of<br />

h<strong>and</strong>edness in the cholesteric phase using two competing chiral<br />

dopants—one of switchable twisting ability,the other one<br />

photoinert—see: T. Sagisaka,Y. Yokoyama,Bull. Chem. Soc.<br />

Jpn. 2000, 73,191 – 196; d) for color control through variation<br />

of pitch in a cholesteric liquid-crystalline film by photoinduced<br />

alteration of the pseudoenantiomeric ratio for an overcrowded<br />

alkene dopant,followed by fixing of the state by photopolymerization,see:<br />

R. A. van Delden,M. B. van Gelder,N. P. M.<br />

Huck,B. L. Feringa,Adv. Funct. Mater. 2003, 13,319 – 324.<br />

[462] R. A. van Delden,N. Koumura,N. Harada,B. L. Feringa,Proc.<br />

Natl. Acad. Sci. USA 2002, 99,4945 – 4949.<br />

[463] D. A. Leigh,M. A. F. Morales,E. M. PØrez,J. K. Y. Wong,C. G.<br />

Saiz,A. M. Z. Slawin,A. J. Carmichael,D. M. Haddleton,<br />

A. M. Brouwer,W. J. Buma,G. W. H. Wurpel,S. León,F.<br />

Zerbetto, Angew. Chem. 2005, 117,3122 – 3127; Angew. Chem.<br />

Int. Ed. 2005, 44,3062 – 3067.<br />

[464] W.-Q. Deng,A. H. Flood,J. F. Stoddart,W. A. Goddard,J. Am.<br />

Chem. Soc. 2005, 127,15 994 – 15995.<br />

[465] For reviews which compare the characteristics of various<br />

different actuator materials,see: a) J. D. W. Madden,N. A.<br />

V<strong>and</strong>esteeg,P. A. Anquetil,P. G. A. Madden,A. Takshi,R. Z.<br />

Pytel,S. R. Lafontaine,P. A. Wieringa,I. W. Hunter,IEEE J.<br />

Oceanic Eng. 2004, 29,706 – 728; b) R. H. Baughman,Science<br />

2005, 308,63 – 65; for reviews concentrating on specific<br />

molecular strategies,see below.<br />

[466] For reviews on the development of this field,<strong>and</strong> progress<br />

towards applications,see: a) T. Tanaka,Sci. Am. 1981, 244(1),<br />

110 – 123; b) H. G. Schild, Prog. Polym. Sci. 1992, 17,163 – 249;<br />

c) K. Kajiwara,S. B. Ross-Murphy,Nature 1992, 355,208 – 209;<br />

d) Y. Osada,J. P. Gong,Prog. Polym. Sci. 1993, 18,187 – 226;<br />

e) Y. Osada,S. B. Ross-Murphy,Sci. Am. 1993, 268(5),82 – 87;<br />

f) Y. Osada,J. P. Gong,Adv. Mater. 1998, 10,827 – 837; g) M.<br />

Shahinpoor,Y. Bar-Cohen,J. O. Simpson,J. Smith,Smart<br />

Mater. Struct. 1998, 7,R15-R30.<br />

[467] a) I. Z. Steinberg,A. Oplatka,A. Katchalsky,Nature 1966, 210,<br />

568 – 571; b) M. V. Sussman,A. Katchalsky,Science 1970, 167,<br />

45 – 47.<br />

[468] Rather than using entirely synthetic gels,in practice Katchalsky<br />

s devices employed cross-linked collagen fibres. The<br />

ability of various natural polypeptides <strong>and</strong> protein bodies to<br />

undergo reversible,stimuli-induced conformational changes<br />

resulting in free-energy transduction has since been extensively<br />

studied,both from the perspective of underst<strong>and</strong>ing such<br />

processes in nature <strong>and</strong> also for artificial applications. For<br />

reviews,see: Ref. [97k] <strong>and</strong> a) D. W. Urry,J. Phys. Chem. B<br />

1997, 101,11007 – 11 028; b) D. W. Urry,Biopolymers 1998, 47,<br />

167 – 178; c) C. Mavroidis,A. Dubey,Nat. Mater. 2003, 2,573 –<br />

574.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

[469] S. Juodkazis,N. Mukai,R. Wakaki,A. Yamaguchi,S. Matsuo,<br />

H. Misawa, Nature 2000, 408,178 – 181.<br />

[470] a) R. Yoshida,T. Takahashi,T. Yamaguchi,H. Ichijo,J. Am.<br />

Chem. Soc. 1996, 118,5134 – 5135; b) R. Yoshida,E. Kokufuta,<br />

T. Yamaguchi, Chaos 1999, 9,260 – 266; c) R. Yoshida,G.<br />

Otoshi,T. Yamaguchi,E. Kokufuta,J. Phys. Chem. A 2001, 105,<br />

3667 – 3672; d) C. J. Crook,A. Smith,R. A. L. Jones,A. J.<br />

Ryan, Phys. Chem. Chem. Phys. 2002, 4,1367 – 1369; e) R.<br />

Yoshida,K. Takei,T. Yamaguchi,Macromolecules 2003, 36,<br />

1759 – 1761; f) J. R. Howse,P. Topham,C. J. Crook,A. J.<br />

Gleeson,W. Bras,R. A. L. Jones,A. J. Ryan,Nano Lett. 2006,<br />

6,73 – 77.<br />

[471] a) C. D. Jones,L. A. Lyon,Macromolecules 2000, 33,8301 –<br />

8306; b) I. Berndt,W. Richtering,Macromolecules 2003, 36,<br />

8780 – 8785; c) I. Berndt,J. S. Pedersen,W. Richtering,J. Am.<br />

Chem. Soc. 2005, 127,9372 – 9373; d) I. Berndt,C. Popescu,F. J.<br />

Wortmann,W. Richtering,Angew. Chem. 2006, 118,1099 –<br />

1102; Angew. Chem. Int. Ed. 2006, 45,1081 – 1085; e) I.<br />

Berndt,J. S. Pedersen,W. Richtering,Angew. Chem. 2006,<br />

118,1769 – 1773; Angew. Chem. Int. Ed. 2006, 45,1737 – 1741.<br />

[472] a) S. Kiyonaka,K. Sugiyasu,S. Shinkai,I. Hamachi,J. Am.<br />

Chem. Soc. 2002, 124,10 954 – 10955; b) S. L. Zhou,S. Matsumoto,H.<br />

D. Tian,H. Yamane,A. Ojida,S. Kiyonaka,I.<br />

Hamachi, Chem. Eur. J. 2005, 11,1130 – 1136.<br />

[473] a) T. Tanaka,C. N. Wang,V. P<strong>and</strong>e,A. Y. Grosberg,A. English,<br />

S. Masamune,H. Gold,R. Levy,K. King,Faraday Discuss.<br />

1995, 101,201 – 206; b) T. Oya,T. Enoki,A. Y. Grosberg,S.<br />

Masamune,T. Sakiyama,Y. Takeoka,K. Tanaka,G. Q. Wang,<br />

Y. Yilmaz,M. S. Feld,R. Dasari,T. Tanaka,Science 1999, 286,<br />

1543 – 1545.<br />

[474] a) D. J. Beebe,J. S. Moore,J. M. Bauer,Q. Yu,R. H. Liu,C.<br />

Devadoss,B. H. Jo,Nature 2000, 404,588 – 590; b) Q. Yu,J. M.<br />

Bauer,J. S. Moore,D. J. Beebe,Appl. Phys. Lett. 2001, 78,<br />

2589 – 2591; c) R. H. Liu,Q. Yu,D. J. Beebe,J. Microelectromech.<br />

Syst. 2002, 11,45 – 53.<br />

[475] Y. Osada,H. Okuzaki,H. Hori,Nature 1992, 355,242 – 244.<br />

[476] a) S. S. Pennadam,M. D. Lavigne,C. F. Dutta,K. Firman,D.<br />

Mernagh,D. C. Górecki,C. Alex<strong>and</strong>er,J. Am. Chem. Soc. 2004,<br />

126,13208 – 13 209; b) S. S. Pennadam,K. Firman,C.<br />

Alex<strong>and</strong>er,D. C. Górecki,J. Nanobiotechnology 2004, 2,8.<br />

[477] For reviews,see: a) I. Y. Galaev,B. Mattiasson,Trends<br />

Biotechnol. 1999, 17,335 – 340; b) N. A. Peppas,Y. Huang,M.<br />

Torres-Lugo,J. H. Ward,J. Zhang,Annu. Rev. Biomed. Eng.<br />

2000, 2,9 – 29; c) Y. Qiu,K. Park,Adv. Drug Delivery Rev. 2001,<br />

53,321 – 339; d) B. Jeong,A. Gutowska,Trends Biotechnol.<br />

2002, 20,305 – 311; e) J. Kopecek,Nature 2002, 417,388 – 391;<br />

f) R. Langer,D. A. Tirrell,Nature 2004, 428,487 – 492.<br />

[478] a) E. Kokufata,Y.-Q. Zhang,T. Tanaka,Nature 1991, 351,302 –<br />

304; b) M. Irie,Y. Misumi,T. Tanaka,Polymer 1993, 34,4531 –<br />

4535; c) J. H. Holtz,S. A. Asher,Nature 1997, 389,829 – 832;<br />

d) J. H. Holtz,J. S. W. Holtz,C. H. Munro,S. A. Asher,Anal.<br />

Chem. 1998, 70,780 – 791; e) T. Miyata,N. Asami,T. Uragami,<br />

Nature 1999, 399,766 – 769; f) K. Lee,S. A. Asher,J. Am.<br />

Chem. Soc. 2000, 122,9534 – 9537; g) C. E. Reese,M. E.<br />

Baltusavich,J. P. Keim,S. A. Asher,Anal. Chem. 2001, 73,<br />

5038 – 5042; h) S. A. Asher,V. L. Alexeev,A. V. Goponenko,<br />

A. C. Sharma,I. K. Lednev,C. S. Wilcox,D. N. Finegold,J. Am.<br />

Chem. Soc. 2003, 125,3322 – 3329; i) S. A. Asher,A. C. Sharma,<br />

A. V. Goponenko,M. M. Ward,Anal. Chem. 2003, 75,1676 –<br />

1683; j) V. L. Alexeev,A. C. Sharma,A. V. Goponenko,S. Das,<br />

I. K. Lednev,C. S. Wilcox,D. N. Finegold,S. A. Asher,Anal.<br />

Chem. 2003, 75,2316 – 2323; k) C. E. Reese,S. A. Asher,Anal.<br />

Chem. 2003, 75,3915 – 3918; l) H.-J. Schneider,T. J. Liu,N.<br />

Lomadze, Angew. Chem. 2003, 115,3668 – 3671; Angew. Chem.<br />

Int. Ed. 2003, 42,3544 – 3546; m) V. L. Alexeev,S. Das,D. N.<br />

Finegold,S. A. Asher,Clin. Chem. 2004, 50,2353 – 2360;<br />

n) A. C. Sharma,T. Jana,R. Kesavamoorthy,L. J. Shi,M. A.<br />

186 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

Virji,D. N. Finegold,S. A. Asher,J. Am. Chem. Soc. 2004, 126,<br />

2971 – 2977; o) H.-J. Schneider,T. J. Liu,Chem. Commun. 2004,<br />

100 – 101; p) H.-J. Schneider,L. Tianjun,N. Lomadze,Chem.<br />

Commun. 2004,2436 – 2437; q) H.-J. Schneider,T. J. Liu,N.<br />

Lomadze,B. Palm,Adv. Mater. 2004, 16,613 – 615; r) A. V.<br />

Goponenko,S. A. Asher,J. Am. Chem. Soc. 2005, 127,10753 –<br />

10759.<br />

[479] For reviews covering shape memory in a number of polymeric<br />

systems,see: a) Z. G. Wei,R. S<strong>and</strong>strom,S. Miyazaki,J. Mater.<br />

Sci. 1998, 33,3743 – 3762; b) A. Lendlein,S. Kelch,Angew.<br />

Chem. 2002, 114,2138 – 2162; Angew. Chem. Int. Ed. 2002, 41,<br />

2034 – 2057.<br />

[480] T. F. Scott,A. D. Schneider,W. D. Cook,C. N. Bowman,<br />

Science 2005, 308,1615 – 1617.<br />

[481] A. Lendlein,H. Y. Jiang,O. Junger,R. Langer,Nature 2005,<br />

434,879 – 882.<br />

[482] For reviews see: a) R. H. Baughman, Synth. Met. 1996, 78,339 –<br />

353; b) T. F. Otero,J. M. Sansiæena,Adv. Mater. 1998, 10,491 –<br />

494; c) E. W. H. Jager,E. Smela,O. Inganas,Science 2000, 290,<br />

1540 – 1545; d) E. Smela, Adv. Mater. 2003, 15,481 – 494.<br />

[483] J. M. Sansiæena,V. Olazµbal,T. F. Otero,C. N. P. da Fonseca,<br />

M.-A. De Paoli, Chem. Commun. 1997,2217 – 2218.<br />

[484] J. D. Madden,R. A. Cush,T. S. Kanigan,C. J. Brenan,I. W.<br />

Hunter, Synth. Met. 1999, 105,61 – 64.<br />

[485] W. Lu,A. G. Fadeev,B. H. Qi,E. Smela,B. R. Mattes,J. Ding,<br />

G. M. Spinks,J. Mazurkiewicz,D. Z. Zhou,G. G. Wallace,D. R.<br />

MacFarlane,S. A. Forsyth,M. Forsyth,Science 2002, 297,983 –<br />

987.<br />

[486] L. M. Low,S. Seetharaman,K. Q. He,M. J. Madou,Sens.<br />

Actuators B 2000, 67,149 – 160.<br />

[487] K. J. Lin,S. J. Fu,C. Y. Cheng,W. H. Chen,H. M. Kao,Angew.<br />

Chem. 2004, 116,4282 – 4285; Angew. Chem. Int. Ed. 2004, 43,<br />

4186 – 4189.<br />

[488] R. H. Baughman,C. X. Cui,A. A. Zakhidov,Z. Iqbal,J. N.<br />

Barisci,G. M. Spinks,G. G. Wallace,A. Mazzoldi,D. De Rossi,<br />

A. G. Rinzler,O. Jaschinski,S. Roth,M. Kertesz,Science 1999,<br />

284,1340 – 1344.<br />

[489] B. Raguse,K.-H. Müller,L. Wieczorek,Adv. Mater. 2003, 15,<br />

922 – 926.<br />

[490] C. D. Eisenbach, Polymer 1980, 21,1175 – 1179.<br />

[491] For extensive reviews of azo-containing polymers <strong>and</strong> their<br />

various switchable properties,see: a) G. S. Kumar,D. C.<br />

Neckers, Chem. Rev. 1989, 89,1915 – 1925; b) S. Xie,A.<br />

Natansohn,P. Rochon,Chem. Mater. 1993, 5,403 – 411; c) A.<br />

Natansohn,P. Rochon,Chem. Rev. 2002, 102,4139 – 4175.<br />

[492] a) T. Hugel,N. B. Holl<strong>and</strong>,A. Cattani,L. Moroder,M. Seitz,<br />

H. E. Gaub, Science 2002, 296,1103 – 1106; b) N. B. Holl<strong>and</strong>,T.<br />

Hugel,G. Neuert,A. Cattani-Scholz,C. Renner,D. Oesterhelt,<br />

L. Moroder,M. Seitz,H. E. Gaub,Macromolecules 2003, 36,<br />

2015 – 2023.<br />

[493] a) Y. Okamoto,T. Nakano,Chem. Rev. 1994, 94,349 – 372; b) T.<br />

Nakano,Y. Okamoto,Chem. Rev. 2001, 101,4013 – 4038.<br />

[494] G. Maxein,R. Zentel,Macromolecules 1995, 28,8438 – 8440.<br />

[495] J. Li,G. B. Schuster,K. S. Cheon,M. M. Green,J. V. Selinger,J.<br />

Am. Chem. Soc. 2000, 122,2603 – 2612.<br />

[496] This process is exactly analogous to the generation of photoinduced<br />

anisotropy in azo-containing liquid crystals (see<br />

Section 8.3.2) <strong>and</strong> is also applicable to polymer systems<br />

containing other photoisomerizable molecules. It is also related<br />

to the “all optical poling” method for ordering non-centrosymmetric<br />

chromophores in polymers. For a discussion of these<br />

mechanims,see: M. Dumont,A. El Osman,Chem. Phys. 1999,<br />

245,437 – 462.<br />

[497] J. A. Delaire,K. Nakatani,Chem. Rev. 2000, 100,1817 – 1845.<br />

[498] a) P. Rochon,E. Batalla,A. Natansohn,Appl. Phys. Lett. 1995,<br />

66,136 – 138; b) D. Y. Kim,S. K. Tripathy,L. Li,J. Kumar,<br />

Appl. Phys. Lett. 1995, 66,1166 – 1168; for accounts of the<br />

Angew<strong>and</strong>te<br />

Chemie<br />

observation of this remarkable effect,see: c) N. K. Viswanathan,D.<br />

Y. Kim,S. P. Bian,J. Williams,W. Liu,L. Li,L.<br />

Samuelson,J. Kumar,S. K. Tripathy,J. Mater. Chem. 1999, 9,<br />

1941 – 1955; d) K. G. Yager,C. J. Barrett,Curr. Opin. Solid<br />

State Mater. Sci. 2001, 6,487 – 494.<br />

[499] The precise mechanism whereby mass transport occurs during<br />

photoorientation is still a topic of intense investigation,see,for<br />

example: Refs. [491c,497,498c,d] <strong>and</strong> a) F. L. Labarthet,J. L.<br />

Bruneel,T. Buffeteau,C. Sourisseau,M. R. Huber,S. J. Zilker,<br />

T. Bieringer, Phys. Chem. Chem. Phys. 2000, 2,5154 – 5167;<br />

b) B. M. Schulz,M. R. Huber,T. Bieringer,G. Krausch,S. J.<br />

Zilker, Synth. Met. 2001, 124,155 – 157.<br />

[500] For a review,see: a) P. Xie,R. Zhang,J. Mater. Chem. 2005, 15,<br />

2529 – 2550; for the first example of a microscopic device in<br />

which this technology is used,see: b) A. Buguin,M. H. Li,P.<br />

Silberzan,B. Ladoux,P. Keller,J. Am. Chem. Soc. 2006, 128,<br />

1088 – 1089.<br />

[501] a) H. Finkelmann,E. Nishikawa,G. G. Pereira,M. Warner,<br />

Phys. Rev. Lett. 2001, 87,015501; b) P. M. Hogan,A. R.<br />

Tajbakhsh,E. M. Terentjev,Phys. Rev. E 2002, 65,041720;<br />

c) J. Cviklinski,A. R. Tajbakhsh,E. M. Terentjev,Eur. Phys. J.<br />

E 2002, 9,427 – 434; d) T. Ikeda,M. Nakano,Y. L. Yu,O.<br />

Tsutsumi,A. Kanazawa,Adv. Mater. 2003, 15,201 – 205;<br />

e) M. H. Li,P. Keller,B. Li,X. G. Wang,M. Brunet,Adv.<br />

Mater. 2003, 15,569 – 572; f) Y. L. Yu,M. Nakano,T. Ikeda,<br />

Nature 2003, 425,145 – 145; g) M. Camacho-Lopez,H. Finkelmann,P.<br />

Palffy-Muhoray,M. Shelley,Nat. Mater. 2004, 3,307 –<br />

310.<br />

[502] It has been proposed that the photoinduced macroscopic<br />

mechanical changes in both azo-containing liquid-crystal elastomers<br />

<strong>and</strong> spin-cast films arise from the same “photomechanical”<br />

effect at the molecular level <strong>and</strong> attention is now turning<br />

to the proper characterization of these phenomena,see: O. M.<br />

Tanchak,C. J. Barrett,Macromolecules 2005, 38,10566 –<br />

10570.<br />

[503] R. Eelkema,M. M. Pollard,J. Vicario,N. Katsonis,B. S.<br />

Ramon,C. W. M. Bastiaansen,D. J. Broer,B. L. Feringa,<br />

Nature 2006, 440,163.<br />

[504] M. Kaneko,K. Kaneto,Synth. Met. 1999, 102,1350 – 1353.<br />

[505] a) M. J. Marsella,R. J. Reid,Macromolecules 1999, 32,5982 –<br />

5984; b) M. J. Marsella,R. J. Reid,S. Estassi,L. S. Wang,J. Am.<br />

Chem. Soc. 2002, 124,12507 – 12 510; c) M. J. Marsella,G. Z.<br />

Piao,F. S. Tham,Synthesis 2002,1133 – 1135; d) M. J. Marsella,<br />

Acc. Chem. Res. 2002, 35,944 – 951.<br />

[506] H. H. Yu,B. Xu,T. M. Swager,J. Am. Chem. Soc. 2003, 125,<br />

1142 – 1143.<br />

[507] a) H. H. Yu,T. M. Swager,IEEE J. Oceanic Eng. 2004, 29,692 –<br />

695; b) D. A. Scherlis,N. Marzari,J. Am. Chem. Soc. 2005, 127,<br />

3207 – 3212; c) J. Casanovas,D. Zanuy,C. Aleman,Angew.<br />

Chem. 2006, 118,1121 – 1123; Angew. Chem. Int. Ed. 2006, 45,<br />

1103 – 1105.<br />

[508] a) T. J. Huang,B. Brough,C.-M. Ho,Y. Liu,A. H. Flood,P. A.<br />

Bonvallet,H.-R. Tseng,J. F. Stoddart,M. Baller,S. Magonov,<br />

Appl. Phys. Lett. 2004, 85,5391 – 5393; b) Y. Liu,A. H. Flood,<br />

P. A. Bonvallet,S. A. Vignon,B. H. Northrop,H.-R. Tseng,<br />

J. O. Jeppesen,T. J. Huang,B. Brough,M. Baller,S. Magonov,<br />

S. D. Solares,W. A. Goddard,C.-M. Ho,J. F. Stoddart,J. Am.<br />

Chem. Soc. 2005, 127,9745 – 9759.<br />

[509] Surface-stress-induced bending of microcantilevers has also<br />

been shown to be an extremely sensitive probe of molecularrecognition<br />

processes occurring at surface-mounted species<br />

(particularly for biomolecules),see for example: a) J. Fritz,<br />

M. K. Baller,H. P. Lang,H. Rothuizen,P. Vettiger,E. Meyer,<br />

H. J. Guntherodt,C. Gerber,J. K. Gimzewski,Science 2000,<br />

288,316 – 318; b) G. H. Wu,H. F. Ji,K. Hansen,T. Thundat,R.<br />

Datar,R. Cote,M. F. Hagan,A. K. Chakraborty,A. Majumdar,<br />

Proc. Natl. Acad. Sci. USA 2001, 98,1560 – 1564; c) G. H. Wu,<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

187


Reviews<br />

R. H. Datar,K. M. Hansen,T. Thundat,R. J. Cote,A.<br />

Majumdar, Nat. Biotechnol. 2001, 19,856 – 860; d) R. McKendry,J.<br />

Y. Zhang,Y. Arntz,T. Strunz,M. Hegner,H. P. Lang,<br />

M. K. Baller,U. Certa,E. Meyer,H. J. Guntherodt,C. Gerber,<br />

Proc. Natl. Acad. Sci. USA 2002, 99,9783 – 9788; e) B. Dhayal,<br />

W. A. Henne,D. D. Doorneweerd,R. G. Reifenberger,P. S.<br />

Low, J. Am. Chem. Soc. 2006, 128,3716 – 3721; mechanical<br />

changes have also been achieved in response to photoisomerization<br />

of an azobenzene monolayer (f) H. F. Ji,Y. Feng,X. H.<br />

Xu,V. Purushotham,T. Thundat,G. M. Brown,Chem.<br />

Commun. 2004,2532 – 2533); <strong>and</strong> pH-switched conformational<br />

changes in an oligonucleotide monolayer (g) W. M. Shu,D. S.<br />

Liu,M. Watari,C. K. Riener,T. Strunz,M. E. Well<strong>and</strong>,S.<br />

Balasubramanian,R. A. McKendry,J. Am. Chem. Soc. 2005,<br />

127,17054 – 17 060); finally,a photoinduced,surface-catalyzed<br />

chemical reaction has been used to reversibly bend a microcantilever<br />

(h) M. Su,V. P. Dravid,Nano Lett. 2005, 5,2023 –<br />

2028).<br />

[510] For a recent review,see: Y. Liu,L. Mu,B. H. Liu,J. L. Kong,<br />

Chem. Eur. J. 2005, 11,2622 – 2631.<br />

[511] a) S. T. Milner, Science 1991, 251,905 – 914; b) V. V. Tsukruk,<br />

Prog. Polym. Sci. 1997, 22,247 – 311; c) T. P. Russell,Science<br />

2002, 297,964 – 967; d) N. Nath,A. Chilkoti,Adv. Mater. 2002,<br />

14,1243 – 1247; e) I. Luzinov,S. Minko,V. V. Tsukruk,Prog.<br />

Polym. Sci. 2004, 29,635 – 698.<br />

[512] V. Pardo-Yissar,R. Gabai,A. N. Shipway,T. Bourenko,I.<br />

Willner, Adv. Mater. 2001, 13,1320 – 1323.<br />

[513] a) G. V. R. Rao,G. P. López,Adv. Mater. 2000, 12,1692 – 1695;<br />

b) G. V. R. Rao,S. Balamurugan,D. E. Meyer,A. Chilkoti,<br />

G. P. López, Langmuir 2002, 18,1819 – 1824; c) G. V. R. Rao,<br />

M. E. Krug,S. Balamurugan,H. F. Xu,Q. Xu,G. P. López,<br />

Chem. Mater. 2002, 14,5075 – 5080.<br />

[514] Control over mass transport through porous materials has<br />

similarly been controlled using photoresponsive (a) N. G. Liu,<br />

D. R. Dunphy,P. Atanassov,S. D. Bunge,Z. Chen,G. P. López,<br />

T. J. Boyle,C. J. Brinker,Nano Lett. 2004, 4,551 – 554) <strong>and</strong> pHresponsive<br />

(b) R. Casasffls,M. D. Marcos,R. Martínez-Mµæez,<br />

J. V. Ros-Lis,J. Soto,L. A. Villaescusa,P. Amorós,D. Beltrµn,<br />

C. Guillem,J. Latorre,J. Am. Chem. Soc. 2004, 126,8612 –<br />

8613) oligomers incorporated into the inorganic matrix.<br />

[515] D. L. Huber,R. P. Manginell,M. A. Samara,B. I. Kim,B. C.<br />

Bunker, Science 2003, 301,352 – 354.<br />

[516] M. Cavallini,R. Lazzaroni,R. Zamboni,F. Biscarini,D.<br />

Timpel,F. Zerbetto,G. J. Clarkson,D. A. Leigh,J. Phys.<br />

Chem. B 2001, 105,10 826 – 10830.<br />

[517] T. D. Nguyen,H.-R. Tseng,P. C. Celestre,A. H. Flood,Y. Liu,<br />

J. F. Stoddart,J. I. Zink,Proc. Natl. Acad. Sci. USA 2005, 102,<br />

10029 – 10034.<br />

[518] a) I. Willner,B. Willner,J. Mater. Chem. 1998, 8,2543 – 2556;<br />

b) V. Pardo-Yissar,E. Katz,I. Willner,A. B. Kotlyar,C.<br />

S<strong>and</strong>ers,H. Lill,Faraday Discuss. 2000,119 – 134; c) A. N.<br />

Shipway,I. Willner,Acc. Chem. Res. 2001, 34,421 – 432.<br />

[519] I. Willner,V. Pardo-Yissar,E. Katz,K. T. Ranjit,J. Electroanal.<br />

Chem. 2001, 497,172 – 177.<br />

[520] E. Katz,L. Sheeney-Haj-Ichia,I. Willner,Angew. Chem. 2004,<br />

116,3354 – 3362; Angew. Chem. Int. Ed. 2004, 43,3292 – 3300.<br />

[521] L. Sheeney-Haj-Ichia,I. Willner,J. Phys. Chem. B 2002, 106,<br />

13094 – 13097.<br />

[522] a) E. Katz,O. Lioubashevsky,I. Willner,J. Am. Chem. Soc.<br />

2004, 126,15520 – 15 532; b) E. Katz,R. Baron,I. Willner,N.<br />

Richke,R. D. Levine,ChemPhysChem 2005, 6,2179 – 2189.<br />

[523] a) R. Blossey, Nat. Mater. 2003, 2,301 – 306; b) T. L. Sun,L.<br />

Feng,X. F. Gao,L. Jiang,Acc. Chem. Res. 2005, 38,644 – 652.<br />

[524] A. J. Pertsin,M. Grunze,H. J. Kreuzer,R. L. C. Wang,Phys.<br />

Chem. Chem. Phys. 2000, 2,1729 – 1733.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

[525] J. Lahann,S. Mitragotri,T. N. Tran,H. Kaido,J. Sundaram,I. S.<br />

Choi,S. Hoffer,G. A. Somorjai,R. Langer,Science 2003, 299,<br />

371 – 374.<br />

[526] X. M. Wang,A. B. Kharitonov,E. Katz,I. Willner,Chem.<br />

Commun. 2003,1542 – 1543.<br />

[527] Y. Liu,L. Mu,B. H. Liu,S. Zhang,P. Y. Yang,J. L. Kong,Chem.<br />

Commun. 2004,1194 – 1195.<br />

[528] a) H. P. Greenspan, J. Fluid Mech. 1978, 84,125 – 143; b) F.<br />

Brochard, Langmuir 1989, 5,432 – 438.<br />

[529] a) T. Ondarçuhu,M. VeyssiØ,J. Phys. II 1991, 1,75 – 85;<br />

b) M. K. Chaudhury,G. M. Whitesides,Science 1992, 256,<br />

1539 – 1541.<br />

[530] The difference in surface free energy leads to different<br />

equilibrium values for the contact angle on opposite sides of<br />

the droplet. The implied asymmetric droplet shape however is<br />

in fact never achieved as it sets up a Laplace pressure gradient<br />

inside the droplet,which rapidly equilibrates to give a<br />

symmetric shape with dynamic contact angles at either side of<br />

roughly equal values,intermediate between the two equilibrium<br />

values. It is this imbalance in contact angles which leads to a<br />

directional force on the droplet so that motion occurs. A<br />

difference in contact angles is the thermodynamic driving force<br />

for the motion,but it is not the sole requirement. A kinetic<br />

requirement is a low contact angle hysteresis (that is,small<br />

differences between receding <strong>and</strong> advancing contact angles for<br />

the surface–liquid system in use). The problem of significant<br />

hysteresis can be surmounted by coupling the system to a<br />

source of r<strong>and</strong>om energy,while using the surface energy<br />

gradient to bias the droplet motion,in a ratchetlike mechanism.<br />

This has been investigated for rapidly condensing droplets from<br />

steam: a) S. Daniel,M. K. Chaudhury,J. C. Chen,Science 2001,<br />

291,633 – 636; <strong>and</strong> for droplets subjected to mechanical<br />

vibrations: b) S. Daniel,S. Sircar,J. Gliem,M. K. Chaudhury,<br />

Langmuir 2004, 20,4085 – 4092.<br />

[531] Earlier,the directional motion of cells over artificial surfaces<br />

with a chemical composition gradient had been observed:<br />

a) S. B. Carter, Nature 1965, 208,1183 – 1187; b) S. B. Carter,<br />

Nature 1967, 213,256 – 260; <strong>and</strong> a brief theoretical explanation<br />

used similar arguments to those developed in more detail later:<br />

c) J. L. Moilliet, Nature 1967, 213,260 – 261.<br />

[532] a) H. Gau,S. Herminghaus,P. Lenz,R. Lipowsky,Science 1999,<br />

283,46 – 49; b) B. S. Gallardo,V. K. Gupta,F. D. Eagerton,L. I.<br />

Jong,V. S. Craig,R. R. Shah,N. L. Abbott,Science 1999, 283,<br />

57 – 60; c) S. Daniel,M. K. Chaudhury,Langmuir 2002, 18,<br />

3404 – 3407; d) H. Suda,S. Yamada,Langmuir 2003, 19,529 –<br />

531; e) S. H. Choi,B. M. Z. Newby,Langmuir 2003, 19,7427 –<br />

7435.<br />

[533] See also self-propelled droplets on homogeneous surfaces<br />

operating by the chemical Marangoni effect (Section 6).<br />

[534] M. Grunze, Science 1999, 283,41 – 42.<br />

[535] a) G. M. Whitesides,A. D. Stroock,Phys. Today 2001, 54(6),<br />

42 – 48; b) D. R. Reyes,D. Iossifidis,P. A. Auroux,A. Manz,<br />

Anal. Chem. 2002, 74,2623 – 2636.<br />

[536] a) K. Ichimura,S. K. Oh,M. Nakagawa,Science 2000, 288,<br />

1624 – 1626; b) S. K. Oh,M. Nakagawa,K. Ichimura,J. Mater.<br />

Chem. 2002, 12,2262 – 2269.<br />

[537] J. Bernµ,D. A. Leigh,M. Lubomska,S. M. Mendoza,E. M.<br />

PØrez,P. Rudolf,G. Teobaldi,F. Zerbetto,Nat. Mater. 2005, 4,<br />

704 – 710.<br />

[538] Motion of an organic droplet along a thin-film metal electrode<br />

immersed in an aqueous electrolyte has very recently been<br />

observed in a system which does not employ significant<br />

submolecular motion to alter the surface free energy,see: R.<br />

Yamada,H. Tada,Langmuir 2005, 21,4254 – 4256.<br />

[539] S. Minko,M. Müller,M. Motornov,M. Nitschke,K. Grundke,<br />

M. Stamm, J. Am. Chem. Soc. 2003, 125,3896 – 3900.<br />

188 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

[540] T. L. Sun,G. J. Wang,L. Feng,B. Q. Liu,Y. M. Ma,L. Jiang,<br />

D. B. Zhu, Angew. Chem. 2004, 116,361 – 364; Angew. Chem.<br />

Int. Ed. 2004, 43,357 – 360.<br />

[541] Q. Fu,G. V. R. Rao,S. B. Basame,D. J. Keller,K. Artyushkova,<br />

J. E. Fulghum,G. P. López,J. Am. Chem. Soc. 2004, 126,8904 –<br />

8905.<br />

[542] R. Rosario,D. Gust,A. A. Garcia,M. Hayes,J. L. Taraci,T.<br />

Clement,J. W. Dailey,S. T. Picraux,J. Phys. Chem. B 2004, 108,<br />

12640 – 12642.<br />

[543] a) H. Noji,R. Yasuda,M. Yoshida,K. Kinosita,Nature 1997,<br />

386,299 – 302; b) R. Yasuda,H. Noji,K. Kinosita,M. Yoshida,<br />

Cell 1998, 93,1117 – 1124; c) K. Adachi,R. Yasuda,H. Noji,H.<br />

Itoh,Y. Harada,M. Yoshida,K. Kinosita,Proc. Natl. Acad. Sci.<br />

USA 2000, 97,7243 – 7247; d) R. Yasuda,H. Noji,M. Yoshida,<br />

K. Kinosita,H. Itoh,Nature 2001, 410,898 – 904.<br />

[544] Y. Sambongi,Y. Iko,M. Tanabe,H. Omote,A. Iwamoto-<br />

Kihara,I. Ueda,T. Yanagida,Y. Wada,M. Futai,Science 1999,<br />

286,1722 – 1724.<br />

[545] H. Itoh,A. Takahashi,K. Adachi,H. Noji,R. Yasuda,M.<br />

Yoshida,K. Kinosita,Nature 2004, 427,465 – 468.<br />

[546] a) C. Montemagno,G. Bach<strong>and</strong>,Nanotechnology 1999, 10,<br />

225 – 231; b) R. K. Soong,G. D. Bach<strong>and</strong>,H. P. Neves,A. G.<br />

Olkhovets,H. G. Craighead,C. D. Montemagno,Science 2000,<br />

290,1555 – 1558.<br />

[547] A number of reviews cover the various aspects of this field in<br />

greater detail than we are able to achieve here,see: a) H. Hess,<br />

V. Vogel, Rev. Mol. Biotechnol. 2001, 82,67 – 85; b) J. J.<br />

Schmidt,C. D. Montemagno in <strong>Molecular</strong> <strong>Motors</strong> (Ed.: M.<br />

Schliwa),Wiley-VCH,Weinheim,2003,pp. 541 – 558; c) H.<br />

Hess,G. D. Bach<strong>and</strong>,V. Vogel,Chem. Eur. J. 2004, 10,2110 –<br />

2116; d) C. Mavroidis,A. Dubey,M. L. Yarmush,Annu. Rev.<br />

Biomed. Eng. 2004, 6,363 – 395; e) M. Knoblauch,W. S. Peters,<br />

Cell. Mol. Life Sci. 2004, 61,2497 – 2509; for a recent review<br />

which places artificial molecular machines side-by-side with<br />

motor proteins,see Ref. [1l].<br />

[548] a) S. J. Kron,J. A. Spudich,Proc. Natl. Acad. Sci. USA 1986, 83,<br />

6272 – 6276; b) Y. Y. Toyoshima,S. J. Kron,E. M. McNally,<br />

K. R. Niebling,C. Toyoshima,J. A. Spudich,Nature 1987, 328,<br />

536 – 539; c) Y. Harada,K. Sakurada,T. Aoki,D. D. Thomas,T.<br />

Yanagida, J. Mol. Biol. 1990, 216,49 – 68.<br />

[549] Unfortunately,such systems are sometimes called “molecular<br />

shuttles”,thus risking confusion with their unrelated rotaxanebased<br />

namesakes (see Sections 4.2–4.3).<br />

[550] a) A. Kakugo,S. Sugimoto,J. P. Gong,Y. Osada,Adv. Mater.<br />

2002, 14,1124 – 1126; b) A. Kakugo,K. Shikinaka,K. Matsumoto,J.<br />

P. Gong,Y. Osada,Bioconjugate Chem. 2003, 14,1185 –<br />

1190; c) A. Kakugo,K. Shikinaka,N. Takekawa,S. Sugimoto,<br />

Y. Osada,J. P. Gong,Biomacromolecules 2005, 6,845 – 849;<br />

d) A. Kakugo,K. Shikinaka,J. P. Gong,Y. Osada,Polymer<br />

2005, 46,7759 – 7770.<br />

[551] J. Z. Xi,J. J. Schmidt,C. D. Montemagno,Nat. Mater. 2005, 4,<br />

180 – 184.<br />

[552] For reviews,see: a) G. A. Schick,A. F. Lawrence,R. R. Birge,<br />

Trends Biotechnol. 1988, 6,159 – 163; b) R. R. Birge,Annu.<br />

Rev. Phys. Chem. 1990, 41,683 – 733; c) D. Oesterhelt,C.<br />

Brauchle,N. Hampp,Q. Rev. Biophys. 1991, 24,425 – 478;<br />

d) Z. P. Chen,R. R. Birge,Trends Biotechnol. 1993, 11,292 –<br />

300; e) N. Hampp,A. Silber,Pure Appl. Chem. 1996, 68,1361 –<br />

1366; f) R. R. Birge,N. B. Gillespie,E. W. Izaguirre,A.<br />

Kusnetzow,A. F. Lawrence,D. Singh,Q. W. Song,E. Schmidt,<br />

J. A. Stuart,S. Seetharaman,K. J. Wise,J. Phys. Chem. B 1999,<br />

103,10746 – 10 766; g) N. Hampp,Chem. Rev. 2000, 100,1755 –<br />

1776.<br />

[553] For reviews,see: Ref. [169d],chapt. 6 <strong>and</strong> a) J.-M. Lehn in<br />

Physical Chemistry of Transmembrane Ion Motions (Ed.: G.<br />

Spach),Elsevier,Amsterdam,1983,pp. 181 – 207; b) G. W.<br />

Gokel,O. Murillo,Acc. Chem. Res. 1996, 29,425 – 432;<br />

Angew<strong>and</strong>te<br />

Chemie<br />

c) T. M. Fyles, Curr. Opin. Chem. Biol. 1997, 1,497 – 505;<br />

d) T. W. Bell, Curr. Opin. Chem. Biol. 1998, 2,711 – 716;<br />

e) G. W. Gokel,A. Mukhopadhyay,Chem. Soc. Rev. 2001, 30,<br />

274 – 286; f) J. M. Boon,B. D. Smith,Curr. Opin. Chem. Biol.<br />

2002, 6,749 – 756; g) B. D. Smith,T. N. Lambert,Chem.<br />

Commun. 2003,2261 – 2268; h) special issue on “<strong>Synthetic</strong><br />

Ion Channels”: Bioorg. Med. Chem. 2004, 12,1277 – 1350; i) S.<br />

Matile,A. Som,N. Sorde,Tetrahedron 2004, 60,6405 – 6435.<br />

[554] S. Sukharev,A. Anishkin,Trends Neurosci. 2004, 27,345 – 351.<br />

[555] a) K. Yoshimura,A. Batiza,M. Schroeder,P. Blount,C. Kung,<br />

Biophys. J. 1999, 77,1960 – 1972; b) K. Yoshimura,A. Batiza,C.<br />

Kung, Biophys. J. 2001, 80,2198 – 2206.<br />

[556] A. Kocer,M. Walko,W. Meijberg,B. L. Feringa,Science 2005,<br />

309,755 – 758.<br />

[557] M. Banghart,K. Borges,E. Isacoff,D. Trauner,R. H. Kramer,<br />

Nat. Neurosci. 2004, 7,1381 – 1386.<br />

[558] M. Volgraf,P. Gorostiza,R. Numano,R. H. Kramer,E. Y.<br />

Isacoff,D. Trauner,Nat. Chem. Biol. 2006, 2,47 – 52.<br />

[559] a) V. Borisenko,D. C. Burns,Z. H. Zhang,G. A. Woolley,J.<br />

Am. Chem. Soc. 2000, 122,6364 – 6370; b) T. Lougheed,V.<br />

Borisenko,T. Hennig,K. Ruck-Braun,G. A. Woolley,Org.<br />

Biomol. Chem. 2004, 2,2798 – 2801.<br />

[560] a) S. S. Anderson,I. G. Lyle,R. Paterson,Nature 1976, 259,<br />

147 – 148; b) J. J. Grimaldi,S. Boileau,J.-M. Lehn,Nature 1977,<br />

265,229 – 230; c) T. Shinbo,K. Kurihara,Y. Kobatake,N.<br />

Kamo, Nature 1977, 270,277 – 278; d) J. J. Grimaldi,J.-M. Lehn,<br />

J. Am. Chem. Soc. 1979, 101,1333 – 1334.<br />

[561] A number of synthetic systems now exist in which relatively<br />

long-lived,long-distance photoinduced intramolecular chargeseparated<br />

states can be generated <strong>and</strong> currently ways of<br />

utilizing this electrical potential are being investigated,for<br />

reviews,see: a) T. J. Meyer,Acc. Chem. Res. 1989, 22,163 – 170;<br />

b) M. R. Wasielewski, Chem. Rev. 1992, 92,435 – 461; c) D.<br />

Gust,T. A. Moore,A. L. Moore,Acc. Chem. Res. 1993, 26,<br />

198 – 205; d) K. Maruyama,A. Osuka,N. Mataga,Pure Appl.<br />

Chem. 1994, 66,867 – 872; e) M. N. Paddon-Row,Acc. Chem.<br />

Res. 1994, 27,18 – 25; f) H. Kurreck,M. Huber,Angew. Chem.<br />

1995, 107,929 – 947; Angew. Chem. Int. Ed. Engl. 1995, 34,849 –<br />

866; g) V. Balzani,A. Credi,M. Venturi,Curr. Opin. Chem.<br />

Biol. 1997, 1,506 – 513; h) D. Gust,T. A. Moore,A. L. Moore,<br />

Acc. Chem. Res. 2001, 34,40 – 48; i) H. Dürr,S. Bossmann,Acc.<br />

Chem. Res. 2001, 34,905 – 917; j) L. C. Sun,L. Hammarstrom,<br />

B. Škermark,S. Styring,Chem. Soc. Rev. 2001, 30,36 – 49; k) D.<br />

Holten,D. F. Bocian,J. S. Lindsey,Acc. Chem. Res. 2002, 35,<br />

57 – 69; l) D. M. Guldi, Chem. Soc. Rev. 2002, 31,22 – 36; m) H.<br />

Imahori,Y. Mori,Y. Matano,J. Photochem. Photobiol. C 2003,<br />

4,51 – 83; n) H. Imahori,Org. Biomol. Chem. 2004, 2,1425 –<br />

1433; o) J. H. Alstrum-Acevedo,M. K. Brennaman,T. J.<br />

Meyer, Inorg. Chem. 2005, 44,6802 – 6827.<br />

[562] G. Steinberg-Yfrach,P. A. Liddell,S. C. Hung,A. L. Moore,D.<br />

Gust,T. A. Moore,Nature 1997, 385,239 – 241.<br />

[563] I. M. Bennett,H. M. V. Farfano,F. Bogani,A. Primak,P. A.<br />

Liddell,L. Otero,L. Sereno,J. J. Silber,A. L. Moore,T. A.<br />

Moore,D. Gust,Nature 2002, 420,398 – 401.<br />

[564] G. Steinberg-Yfrach,J. L. Rigaud,E. N. Durantini,A. L.<br />

Moore,D. Gust,T. A. Moore,Nature 1998, 392,479 – 482.<br />

[565] Z. C. He,S. B. Colbran,D. C. Craig,Chem. Eur. J. 2003, 9,116 –<br />

129.<br />

[566] For reviews,see: a) C. M. Niemeyer,Angew. Chem. 1997, 109,<br />

603 – 606; Angew. Chem. Int. Ed. Engl. 1997, 36,585 – 587;<br />

b) N. C. Seeman, Angew. Chem. 1998, 110,3408 – 3428; Angew.<br />

Chem. Int. Ed. 1998, 37,3220 – 3238; c) N. C. Seeman,Annu.<br />

Rev. Biophys. Biomol. Struct. 1998, 27,225 – 248; d) N. C.<br />

Seeman,H. Wang,X. P. Yang,F. R. Liu,C. D. Mao,W. Q. Sun,<br />

L. Wenzler,Z. Y. Shen,R. J. Sha,H. Yan,M. H. Wong,P. Sa-<br />

Ardyen,B. Liu,H. X. Qiu,X. J. Li,J. Qi,S. M. Du,Y. W.<br />

Zhang,J. E. Mueller,T. J. Fu,Y. L. Wang,J. H. Chen,Nano-<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

189


Reviews<br />

technology 1998, 9,257 – 273; e) N. C. Seeman,Trends Biotechnol.<br />

1999, 17,437 – 443; f) C. M. Niemeyer,Appl. Phys. A 1999,<br />

68,119 – 124; g) C. M. Niemeyer,Curr. Opin. Chem. Biol. 2000,<br />

4,609 – 618; h) N. C. Seeman,A. M. Belcher,Proc. Natl. Acad.<br />

Sci. USA 2002, 99,6451 – 6455; i) N. C. Seeman,Nature 2003,<br />

421,427 – 431; j) N. C. Seeman,Biochemistry 2003, 42,7259 –<br />

7269; k) N. C. Seeman, Chem. Biol. 2003, 10,1151 – 1159;<br />

l) N. C. Seeman, Sci. Am. 2004, 290(6),34 – 43; m) H. Yan,<br />

Science 2004, 306,2048 – 2049.<br />

[567] For reviews of DNA-based machines,see: a) C. M. Niemeyer,<br />

M. Adler, Angew. Chem. 2002, 114,3933 – 3937; Angew. Chem.<br />

Int. Ed. 2002, 41,3779 – 3783; b) N. C. Seeman,Trends Biochem.<br />

Sci. 2005, 30,119 – 125; c) F. C. Simmel,W. U. Dittmer,<br />

Small 2005, 1,284 – 299.<br />

[568] X. P. Yang,A. V. Vologodskii,B. Liu,B. Kemper,N. C. Seeman,<br />

Biopolymers 1998, 45,69 – 83.<br />

[569] C. D. Mao,W. Q. Sun,Z. Y. Shen,N. C. Seeman,Nature 1999,<br />

397,144 – 146.<br />

[570] B. Yurke,A. J. Turberfield,A. P. Mills,F. C. Simmel,J. L.<br />

Neumann, Nature 2000, 406,605 – 608.<br />

[571] F. C. Simmel,B. Yurke,Phys. Rev. E 2001, 6304,041913.<br />

[572] F. C. Simmel,B. Yurke,Appl. Phys. Lett. 2002, 80,883 – 885.<br />

[573] H. Yan,X. P. Zhang,Z. Y. Shen,N. C. Seeman,Nature 2002,<br />

415,62 – 65.<br />

[574] L. P. Feng,S. H. Park,J. H. Reif,H. Yan,Angew. Chem. 2003,<br />

115,4478 – 4482; Angew. Chem. Int. Ed. 2003, 42,4342 – 4346.<br />

[575] D. S. Liu,S. Balasubramanian,Angew. Chem. 2003, 115,5912 –<br />

5914; Angew. Chem. Int. Ed. 2003, 42,5734 – 5736.<br />

[576] This process has been monitored for surface-mounted arrays of<br />

DNA str<strong>and</strong>s by fluorescence switching when a dye is attached<br />

(D. S. Liu,A. Bruckbauer,C. Abell,S. Balasubramanian,D. J.<br />

Kang,D. Klenerman,D. J. Zhou,J. Am. Chem. Soc. 2006, 128,<br />

2067 – 2071) <strong>and</strong> by mechanical changes when the str<strong>and</strong>s are<br />

mounted on microcantilever beams (Ref. [509g]).<br />

[577] a) Y. Chen,S. H. Lee,C. Mao,Angew. Chem. 2004, 116,5449 –<br />

5452; Angew. Chem. Int. Ed. 2004, 43,5335 – 5338; b) M.<br />

Brucale,G. Zuccheri,B. Samori,Org. Biomol. Chem. 2005, 3,<br />

575 – 577.<br />

[578] a) J. W. J. Li,W. H. Tan,Nano Lett. 2002, 2,315 – 318; b) P.<br />

Alberti,J. L. Mergny,Proc. Natl. Acad. Sci. USA 2003, 100,<br />

1569 – 1573.<br />

[579] W. U. Dittmer,A. Reuter,F. C. Simmel,Angew. Chem. 2004,<br />

116,3634 – 3637; Angew. Chem. Int. Ed. 2004, 43,3550 – 3553.<br />

[580] Y. Chen,M. Wang,C. Mao,Angew. Chem. 2004, 116,3638 –<br />

3641; Angew. Chem. Int. Ed. 2004, 43,3554 – 3557.<br />

[581] Y. Chen,C. Mao,J. Am. Chem. Soc. 2004, 126,8626 – 8627.<br />

[582] A. J. Turberfield,J. C. Mitchell,B. Yurke,A. P. Mills,M. I.<br />

Blakey,F. C. Simmel,Phys. Rev. Lett. 2003, 90,118102.<br />

[583] For an overview of this series of publications,see: T. R. Kelly,<br />

Angew. Chem. 2005, 117,4194 – 4198; Angew. Chem. Int. Ed.<br />

2005, 44,4124 – 4127.<br />

[584] W. B. Sherman,N. C. Seeman,Nano Lett. 2004, 4,1203 – 1207.<br />

[585] J. S. Shin,N. A. Pierce,J. Am. Chem. Soc. 2004, 126,10834 –<br />

10835.<br />

[586] T. Ye,C. Mao,J. Am. Chem. Soc. 2004, 126,11410 – 11 411.<br />

[587] a) J. H. Reif in DNA Computing: 8th International Workshop<br />

on DNA-based Computers: Lecture Notes in Computer Science,<br />

Vol. 2568 (Eds.: M. Hagiya,A. Ohuchi),Springer,Berlin,2003,<br />

pp. 22 – 37; b) P. Yin,A. J. Turberfield,J. H. Reif in DNA<br />

Computing: 10th International Workshop on DNA Computing:<br />

Lecture Notes in Computer Science, Vol. 3384 (Eds.: C. Ferretti,<br />

G. Mauri,C. Z<strong>and</strong>ron),Springer,Berlin,2005,pp. 410 – 425.<br />

[588] P. Yin,H. Yan,X. G. Daniell,A. J. Turberfield,J. H. Reif,<br />

Angew. Chem. 2004, 116,5014 – 5019; Angew. Chem. Int. Ed.<br />

2004, 43,4906 – 4911.<br />

[589] J. Bath,S. J. Green,A. J. Turberfield,Angew. Chem. 2005, 117,<br />

4432 – 4435; Angew. Chem. Int. Ed. 2005, 44,4358 – 4361.<br />

D. A. Leigh, F. Zerbetto, <strong>and</strong> E. R. Kay<br />

[590] Y. Tian,Y. He,Y. Chen,P. Yin,C. Mao,Angew. Chem. 2005,<br />

117,4429 – 4432; Angew. Chem. Int. Ed. 2005, 44,4355 – 4358.<br />

[591] Although related to an information ratchet,mechanisms for<br />

Brownian particle motion based on particle-induced irreversible<br />

changes to a symmetrical,periodic potential-energy<br />

surface have been studied in their own right <strong>and</strong> have been<br />

termed “burnt-bridges” mechanisms; see Section 4.4 <strong>and</strong>: a) J.<br />

Mai,I. M. Sokolov,A. Blumen,Phys. Rev. E 2001, 64,011102;<br />

b) T. Antal,P. L. Krapivsky,Phys. Rev. E 2005, 72,046104.<br />

[592] B. Samori,G. Zuccheri,Angew. Chem. 2005, 117,1190 – 1206;<br />

Angew. Chem. Int. Ed. 2005, 44,1166 – 1181.<br />

[593] a) W. U. Dittmer,F. C. Simmel,Nano Lett. 2004, 4,689 – 691;<br />

b) W. U. Dittmer,S. Kempter,J. O. Rädler,F. C. Simmel,Small<br />

2005, 1,708 – 712<br />

[594] Y. Chen,C. Mao,J. Am. Chem. Soc. 2004, 126,13240 – 13241.<br />

[595] a) J. J. Storhoff,C. A. Mirkin,Chem. Rev. 1999, 99,1849 – 1862;<br />

b) E. Braun,K. Keren,Adv. Phys. 2004, 53,441 – 496; c) J.<br />

Wengel, Org. Biomol. Chem. 2004, 2,277 – 280; d) U. Feldkamp,C.<br />

M. Niemeyer,Angew. Chem. 2006, 118,1888 – 1910;<br />

Angew. Chem. Int. Ed. 2006, 45,1856 – 1876.<br />

[596] a) N. E. Broude, Trends Biotechnol. 2002, 20,249 – 256;<br />

b) W. H. Tan,K. M. Wang,T. J. Drake,Curr. Opin. Chem.<br />

Biol. 2004, 8,547 – 553.<br />

[597] W. Q. Shen,M. F. Bruist,S. D. Goodman,N. C. Seeman,<br />

Angew. Chem. 2004, 116,4854 – 4856; Angew. Chem. Int. Ed.<br />

2004, 43,4750 – 4752.<br />

[598] R. Nutiu,Y. F. Li,Angew. Chem. 2005, 117,5600 – 5603; Angew.<br />

Chem. Int. Ed. 2005, 44,5464 – 5467.<br />

[599] For highlights <strong>and</strong> reviews,see: a) A. Gibbons,M. Amos,D.<br />

Hodgson, Curr. Opin. Biotechnol. 1997, 8,103 – 106; b) J. C.<br />

Cox,D. S. Cohen,A. D. Ellington,Trends Biotechnol. 1999, 17,<br />

151 – 154; c) A. J. Ruben,L. F. L<strong>and</strong>weber,Nat. Rev. Mol. Cell<br />

Biol. 2000, 1,69 – 72; d) S. D. Gillmor,P. P. Rugheimer,M. G.<br />

Lagally, Surf. Sci. 2002, 500,699 – 721; e) J. H. Reif,Science<br />

2002, 296,478 – 479; f) M. S. Livstone,D. van Noort,L. F.<br />

L<strong>and</strong>weber, Trends Biotechnol. 2003, 21,98 – 101; g) A.<br />

Condon, Nature 2004, 429,351 – 352.<br />

[600] For reviews,see: a) J. Hasty,D. McMillen,J. J. Collins,Nature<br />

2002, 420,224 – 230; b) R. Weiss,S. Basu,S. Hooshangi,A.<br />

Kalmbach,D. Karig,R. Mehreja,I. Netravali,Nat. Comput.<br />

2003, 2,47 – 84.<br />

[601] M. Freemantle, Chem. Eng. News 2005, 83,33 – 35.<br />

[602] a) P. Hamm,M. Lim,W. F. DeGrado,R. M. Hochstrasser,Proc.<br />

Natl. Acad. Sci. USA 1999, 96,2036 – 2041; b) J. Bredenbeck,J.<br />

Helbing,A. Sieg,T. Schrader,W. Zinth,C. Renner,R.<br />

Behrendt,L. Moroder,J. Wachtveitl,P. Hamm,Proc. Natl.<br />

Acad. Sci. USA 2003, 100,6452 – 6457; c) J. Bredenbeck,J.<br />

Helbing,P. Hamm,J. Am. Chem. Soc. 2004, 126,990 – 991;<br />

d) M. L. Cowan,B. D. Bruner,N. Huse,J. R. Dwyer,B. Chugh,<br />

E. T. J. Nibbering,T. Elsaesser,R. J. D. Miller,Nature 2005,<br />

434,199 – 202; e) J. Zheng,K. Kwak,J. Ashbury,X. Chen,I. R.<br />

Piletic,M. D. Fayer,Science 2005, 309,1338 – 1343; f) H. S.<br />

Chung,M. Khalil,A. W. Smith,Z. Ganim,A. Tokmakoff,Proc.<br />

Natl. Acad. Sci. USA 2005, 102,612 – 617; g) O. F. A. Larsen,P.<br />

Bodis,W. J. Buma,J. S. Hannam,D. A. Leigh,S. Woutersen,<br />

Proc. Natl. Acad. Sci. USA 2005, 102,13378 – 13 382.<br />

[603] V. A. Azov,A. Beeby,M. Cacciarini,A. G. Cheetham,F.<br />

Diederich,M. Frei,J. K. Gimzewski,V. Gramlich,B. Hecht,B.<br />

Jaun,T. Latychevskaia,A. Lieb,Y. Lill,F. Marotti,A. Schlegel,<br />

R. R. Schlittler,P. J. Skinner,P. Seiler,Y. Yamakoshi,Adv.<br />

Funct. Mater. 2006, 16,147 – 156.<br />

[604] J. Clayden,N. Vassiliou,Org. Biomol. Chem. 2006, 4,2667 –<br />

2678.<br />

[605] D. A. Leigh,E. M. PØrez,Top. Curr. Chem. 2006, 265,185 – 208.<br />

[606] A. Credi, Aust. J. Chem. 2006, 59,157 – 169.<br />

[607] H. Tian,Q.-C. Wang,Chem. Soc. Rev. 2006, 35,361 – 374.<br />

190 www.angew<strong>and</strong>te.org 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2007, 46, 72 – 191


<strong>Molecular</strong> Devices<br />

[608] S. Bonnet,J.-P. Collin,M. Koizumi,P. Mobian,J.-P. Sauvage,<br />

Adv. Mater. 2006, 18,1239 – 1250.<br />

[609] T.-A. V. Khuong,J. E. Nuæez,C. E. Godinez,M. A. Garcia-<br />

Garibay, Acc. Chem. Res. 2006, 39,413 – 422.<br />

[610] W. F. Paxton,S. Sundararajan,T. E. Mallouk,A. Sen,Angew.<br />

Chem. 2006, 118,5546 – 5556; Angew. Chem. Int. Ed. 2006, 45,<br />

5420 – 5429.<br />

[611] R. Beckman,K. Beverly. A. Boukai,Y. Bunimovich,J. W. Choi,<br />

E. DeIonno,J. Green,E. Johnston-Halperin,Y. Luo,B. Sheriff,<br />

J. F. Stoddart,J. R. Heath,Faraday Discuss. 2006, 131,9 – 22.<br />

[612] J. Hoogboom,T. Rasing,A. E. Rowan,R. J. M. Nolte,J. Mater.<br />

Chem. 2006, 16,1305 – 1314.<br />

[613] Y. Yu,T. Ikeda,Angew. Chem. 2006, 118,5542 – 5544; Angew.<br />

Chem. Int. Ed. 2006, 45,5416 – 5418.<br />

[614] H. Hess, Soft Matter 2006, 2,669 – 677.<br />

[615] C. C. de Souza Silva,J. Van de Vondel,M. Morelle,V. V.<br />

Moshchalkov, Nature 2006, 440,651 – 654.<br />

[616] a) R. Romeo,S. Carnabuci,L. Fenech,M. Rosaria Plutino,A.<br />

Albinati, Angew. Chem. 2006, 118,4606 – 4610; Angew. Chem.<br />

Int. Ed. 2006, 45,4494 – 4498; b) T. Benincori,G. Celentano,T.<br />

Pilati,A. Ponti,S. Rizzo,F. Sannicolò,Angew. Chem. 2006, 118,<br />

6339 – 6342; Angew. Chem. Int. Ed. 2006, 45,6193 – 6196;<br />

c) R. E. Bulo,F. Allaart,A. W. Ehlers,F. J. J. de Kanter,M.<br />

Schakel,M. Lutz,A. L. Spek,K. Lammertsma,J. Am. Chem.<br />

Soc. 2006, 128,12169 – 12 173.<br />

[617] a) A. J. Nawara,T. Shima,F. Hampel,J. A. Gladysz,J. Am.<br />

Chem. Soc. 2006, 128,4962 – 4963; b) L. Wang,F. Hampel,J. A.<br />

Gladysz, Angew. Chem. 2006, 118,4479 – 4482; Angew. Chem.<br />

Int. Ed. 2006, 45,4372 – 4375; c) S. D. Karlen,C. E. Godinez,<br />

M. A. Garcia-Garibay, Org. Lett. 2006, 8,3417 – 3420.<br />

[618] O. P. H. Vaughan,F. J. Williams,N. Bampos,R. M. Lambert,<br />

Angew. Chem. 2006, 118,3863 – 3865; Angew. Chem. Int. Ed.<br />

2006, 45,3779 – 3781.<br />

[619] W. R. Browne,M. M. Pollard,B. de Lange,A. Meetsma,B. L.<br />

Feringa, J. Am. Chem. Soc. 2006, 128,12412 – 12413.<br />

[620] J. Vicario,M. Walko,A. Meetsma,B. L. Feringa,J. Am. Chem.<br />

Soc. 2006, 128,5127 – 5135.<br />

[621] a) T. Muraoka,K. Kinbara,T. Aida,Nature 2006, 440,512 –<br />

515; for a highlight of this work,see: b) F. M. Raymo,Angew.<br />

Chem. 2006, 118,5375 – 5377; Angew. Chem. Int. Ed. 2006, 45,<br />

5249 – 5251.<br />

[622] T. Muraoka,K. Kinbara,T. Aida,J. Am. Chem. Soc. 2006, 128,<br />

11600 – 11605.<br />

[623] J.-F. Morin,Y. Shirai,J. M. Tour,Org. Lett. 2006, 8,1713 – 1716.<br />

[624] B. Brough,B. H. Northrop,J. J. Schmidt,H.-R. Tseng,K. N.<br />

Houk,J. F. Stoddart,C.-M. Ho,Proc. Natl. Acad. Sci. USA<br />

2006, 103,8583 – 8588.<br />

Angew<strong>and</strong>te<br />

Chemie<br />

[625] P. M. Mendes,W. Lu,H.-R. Tseng,S. Shinder,T. Iijima,M.<br />

Miyaji,C. M. Knobler,J. F. Stoddart,J. Phys. Chem. B 2006,<br />

110,3845 – 3848.<br />

[626] E. DeIonno,H.-R. Tseng,D. D. Harvey,J. F. Stoddart,J. R.<br />

Heath, J. Phys. Chem. B 2006, 110,7609 – 7612.<br />

[627] a) A. H. Flood,E. W. Wong,J. F. Stoddart,Chem. Phys. 2006,<br />

324,280-290; b) A. H. Flood,S. Nygaard,B. W. Laursen,J. O.<br />

Jeppesen,J. F. Stoddart,Org. Lett. 2006, 8,2205 – 2208.<br />

[628] T. J. Huang,A. H. Flood,B. Brough,Y. Liu,P. A. Bonvallet,S.<br />

Kang,C.-W. Chu,T.-F. Guo,W. Lu,Y. Yang,J. F. Stoddart,C.-<br />

M. Ho, IEEE Trans. Autom. Sci. Eng. 2006, 3,254 – 259.<br />

[629] H. Kawai,T. Umehara,K. Fujiwara,T. Tsuji,T. Suzuki,Angew.<br />

Chem. 2006, 118,4387 – 4392; Angew. Chem. Int. Ed. 2006, 45,<br />

4281 – 4286.<br />

[630] K. Jensen,Ç. Girit,W. Mickelson,A. Zettl,Phys. Rev. Lett.<br />

2006, 96,215503.<br />

[631] V. V. Deshp<strong>and</strong>e,H.-Y. Chiu,H. W. Ch. Postma,C. Milkó,L.<br />

Forró,M. Bockrath,Nano Lett. 2006, 6,1092 – 1095.<br />

[632] C. Li,J. Madsen,S. P. Armes,A. L. Lewis,Angew. Chem. 2006,<br />

118,3590 – 3593; Angew. Chem. Int. Ed. 2006, 45,3510 – 3513.<br />

[633] G. K. Samoei,W. Wang,J. O. Escobedo,X. Xu,H.-J. Schneider,<br />

R. L. Cook,R. M. Strongin,Angew. Chem. 2006, 118,5445 –<br />

5448; Angew. Chem. Int. Ed. 2006, 45,5319 – 5322.<br />

[634] L. Dong,A. K. Agarwal,D. J. Beebe,H. Jiang,Nature 2006,<br />

442,551 – 554.<br />

[635] a) V. H. Ebron,Z. Yang,D. J. Seyer,M. E. Kozlov,J. Oh,H.<br />

Xie,J. Razal,L. J. Hall,J. P. Ferraris,A. J. MacDiarmid,R. H.<br />

Baughman, Science 2006, 311,1580 – 1583; for a highlight of this<br />

work,see: b) J. D. Madden,Science 2006, 311,1559 – 1560.<br />

[636] V. Iancu,S.-W. Hla,Proc. Natl. Acad. Sci. USA 2006, 103,<br />

13718 – 13721.<br />

[637] A. Koçer,M. Walko,E. Bulten,E. Halza,B. L. Feringa,W.<br />

Meijberg, Angew. Chem. 2006, 118,3198 – 3202; Angew. Chem.<br />

Int. Ed. 2006, 45,3126 – 3130.<br />

[638] T. Kiwada,K. Sonomura,Y. Sugiura,K. Asami,S. Futaki,J.<br />

Am. Chem. Soc. 2006, 128,6010 – 6011.<br />

[639] a) M. G. L. van den Heuvel,M. P. de Graaff,C. Dekker,Science<br />

2006, 312,910 – 914; for a highlight of this work,see: b) H.<br />

Hess, Science 2006, 312,860 – 861.<br />

[640] T. Liedl,M. Olapinski,F. C. Simmel,Angew. Chem. 2006, 118,<br />

5129 – 5132; Angew. Chem. Int. Ed. 2006, 45,5007 – 5010.<br />

[641] a) S. Bhosale,A. L. Sisson,P. Talukdar,A. Fürstenberg,N.<br />

Banerji,E. Vauthey,G. Bollot,J. Mareda,C. Röger,F.<br />

Würthner,N. Sakai,S. Matile,Science 2006, 313,84 – 86; for a<br />

highlight of this work,see: b) K. Kinbara,T. Aida,Science 2006,<br />

313,51 – 52.<br />

Angew. Chem. Int. Ed. 2007, 46, 72 – 191 2007 Wiley-VCHVerlag GmbH& Co. KGaA, Weinheim www.angew<strong>and</strong>te.org<br />

191

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!